<<

arXiv:hep-th/9206084v1 23 Jun 1992 l fhih mle hn2 than smaller height of all includ they characterized; succinctly are vertices string atreuto n hscnitn RTqatzto fth of imp BRST structure consistent thus algebraic and genus equation master higher The theory. classical tigdarm r ufcswt iia ramtis fol metrics, area minimal 2 with surfaces are diagrams string ler,adteB qaino h ffselsaesaew de we space state off-shell the on equation BV the and algebra, Veuto npyia ttsta eercnl construct recently were that states physical on equation BV tigfil ler stehmtp i algebra Lie strin homotopy of the chain is infinite algebra an field for string identities Jacobi-like to sati vertices rise elementary by defined is nonpolynomial π hs erc eeaieqartcdffrnil nta f that in differentials quadratic generalize metrics These . ⋆ h opeeqatmter fcvratcoe tig sc is strings closed covariant of theory complete The yDOE otatD-C27E009adNFgatPHY91-0 grant NSF Ca and MIT, DE-AC02-76ER03069 Physics, contract Theoretical D.O.E. by for Center address: Permanent LSDSRN IL HOY UNU ACTION QUANTUM THEORY: STRING CLOSED N H - ATREQUATION MASTER B-V THE AND π . nttt o dacdStudy Advanced for Institute colo aua Sciences Natural of School atnZwiebach Barton rneo,N 08540 NJ Princeton, ABSTRACT le Lane Olden L ∞ h ufcswoeflainbnsare bands foliation whose surfaces the e noigteguesmer fthe of symmetry gauge the encoding ⋆ big,Ms.019 upre npart in Supported 02139. Mass. mbridge, edpout.Tegnszero genus The products. field g figrcrinrltosta give that relations recursion sfying ae ycoe edsc flength of geodesics closed by iated isteBtlnVloik (BV) Batalin-Vilkovisky the lies unu cin rmthe From action. quantum e din ed 6210. lainbnscncos The cross. can bands oliation iethe rive ntutdi eal The detail. in onstructed d tigter.The theory. string 2 = ue1992 June hep-th/9206084 MIT-CTP-2102 IASSNS-HEP-92/41 L ∞ ler,adthe and algebra, L ∞ 1. Introduction and Summary of Results

One can hardly overstate the need for a complete formulation of . This necessity arises both at a conceptual level and at a computational level. At the conceptual level we do not yet know the underlying principles of string theory. Such understanding appears to be a prerequisite for addressing deep issues of quantum gravity in the context of string theory. At the computational level we do not have yet the ability to calculate nonperturbative effects. Such effects seem to be needed in order to obtain realistic string models. The subject of this paper is the covariant theory of closed string fields [1–8] This theory, I believe, represents concrete progress in the construction of a complete formulation of string theory. The current formulation of closed string field theory is based in the BRST approach, which originated in the work of Siegel [9]. The theory appears to be the natural closed string analog of the interacting open string field theory constructed by Witten [10]. As we will discuss, it succeeds in generating the perturbative definition of the theory starting from an action based on a gauge principle. It also makes conceptually manifest the factorization and unitarity of the string amplitudes. The closed string field theory is apparently the first field theory for which the most sophisticated machinery for quantization, the Batalin-Vilkovisky (BV) field-antifield formalism [11–13], is necessary and useful in its full form. At present, however, the closed string field theory is not yet a complete formulation of string theory. Its most glaring shortcoming is that its formulation requires making a choice of a conformal background. In a complete formulation, this background should arise as a classical solution. Therefore, the question whether string field theory is the ‘right’ approach to string theory, is at present the question whether string field theory can be formulated without having to choose a conformal background. We may be encouraged by the fact that despite initial difficulties the construction of covariant closed string field theory was possible. The resulting theory, as will be elaborated here, has an interesting and novel algebraic structure arising from subtle properties of moduli spaces of Riemann surfaces. Moreover, as recent work of Sen indicates, the string field theory, while not manifestly independent of the conformal background chosen to formulate it, does incorporate such background independence to some degree [14]. There is therefore little evidence that we face difficulties that cannot altogether be solved in the framework of string field theory by extensions or modifications of the current formulations.

The real obstruction to the construction of closed string field theory was, all along, the absence of a decomposition of the moduli spaces of Riemann surfaces compatible with covariant Feynman rules. A decomposition suitable for the classical theory (genus zero surfaces) was found in Refs.[1,2] by incorporating and generalizing the closed string extension of the open string vertex [10] and the tetrahedron of Ref. [4]. The mathematically precise proof that this was a correct decomposition revealed that the string diagrams were given by minimal area metrics [8], and this suggested a natural way to generalize the construction for higher genus surfaces. The decomposition of moduli space for the quantum theory was addressed in Ref.[5] using minimal area metrics and the ideas of [7], and the algebraic structure of the resulting theory was sketched in Refs.[5,6]. The assumptions used in these works regarding plumbing

2 properties of minimal area metrics have now been established to a large degree in recent works by Ranganathan, Wolf, and the author [15,16]. Even though the main difficulty was geometrical, working out the algebraic implications of the geometrical constructions is both instructive and necessary. This amounts to giving the explicit construction of the string field vertices, with due care of technical issues like antighost insertions, and proving the set of identities they satisfy. These properties must also be shown to guarantee the consistency of the quantum theory constructed with the string field vertices. This is one of the main purposes of the present paper, it amounts to giving the detailed derivation and construction of the quantum closed string field theory sketched in Refs.[5,6]. Essentially we do for the full quantum theory what Kugo and Suehiro, in their important work, did for the classical theory [3]. We will also present new results involving an improved understanding of the algebraic structure of the theory, as well as new results on the concrete specification of the string vertices and the characteristics of minimal area metrics. Furthermore, we will show how to derive the algebraic structures on physical states recently discovered [17,18] in the context of two-dimensional string theory from the corresponding generalized structures of closed string field theory that are defined on the full off-shell state space. Let us consider some of the main points in the present paper.

Algebraic Structure of Closed . The algebraic ingredients of the classical closed string field theory, as emphasized in [19], are a set of multilinear, graded-commutative string products m , with n 2, (n is the number of string fields to be multiplied) satisfying n ≥ a set of Jacobi-type identities involving the BRST operator Q. This structure was seen to correspond to a homotopy [20]. This homotopy Lie algebra will be denoted as L∞, in analogy to the homotopy associative A∞ algebras [21], which are relavant to open string field theory and Chern-Simons theory [22,23]. We identify the closed string L∞ structure precisely by choosing a set of conventions that eliminates the need for insertions in the identities ⋆ relating the string products. This allows to interpret the BRST operator as generating the first product m of the homotopy Lie algebra ( B QB where B is a string field), which 1 ≡ streamlines the proof of gauge invariance. We compute  the algebra of gauge transformations, and determine the field dependent structure constants and the extra terms that vanish upon use of the equations of motion. At the quantum level we have string products labeled by genus. The identity satisfied by all the products is given. This structure is not anymore an L∞ structure due to the appearance of new terms involving a ‘trace’ over two of the entries of the string products. This new structure is shown to guarantee that the master action satisfies the quantum BV master equation. To this end, following the earlier work in open strings [24–26], we give a convenient assignment of fields and antifields in closed string field theory. Extending a result of Sen [27], we show that upon shifting the string field by a term that does not satisfy the classical field equations, the L∞ algebra of the classical theory involving products mn with m 1, acquires the zeroth order product m , this product being just a special state vector. ≥ 0 Despite the indirect construction, this is the algebraic structure of closed string field theory

⋆ I thank E. Witten for suggesting such change of conventions.

3 formulated around a background that is not conformal. If we denote the special state by F , then the lowest identities of the resulting L∞ algebra imply that the generalized Q satisfies QF = 0, and Q2B = [F, B] for all string fields B. The new Q also fails to be a derivation − with respect to the product, the failure being related to F being nonzero. We suggest the possible relevance of these structures to the problem of formulating string field theory without use of conformal backgrounds.

Algebraic Structures on Physical States. The L∞ structure and the BV structure of the closed string field theory is defined on the complete off-shell state space of the theory. On the other hand in recent works similar structures were obtained on the physical states of two-dimensional string theory [17,18]. The homotopy Lie algebra on physical states has m0 = m1 = 0, and the first nonvanishing product is m2. Here we show that, as conjectured in these works, the algebraic structures on physical states can be derived from those on the off-shell state space. This is general, and not peculiar to D = 2 string theory. We do this by showing how to truncate the basic identity of string field theory to physical states. This cannot be done naively with the original string products because of two reasons. First, when the closed string field products mn (for m 3) are used to multiply physical fields, the resulting product is a physical field plus a ≥ BRST-trivial field, plus an unphysical field, which is problematic since it cannot be absorbed as part of the physical fields. Thus the original products do not close on physical fields. A second difficulty is that the original products, since they involve subsets of moduli spaces, would not agree with the physical products, which were defined by correlators over the full moduli spaces. These two difficulties have the same solution. We take a one parameter family of string field products, as the parameter grows the subsets defining the products grow, until they eventually cover the full moduli spaces. In this limit we can see that BRST decoupling of trivial states implies that the product of physical states cannot give an unphysical state, and we are able to derive the identities on physical states from the main identity of string field theory. For the full master equation this requires the use of the Kugo-Ojima quartet mechanism [28].

From Riemann surfaces to Feynman diagrams. The main difficulty in constructing string field theory was due to the lack of a suitable decomposition of the moduli spaces of closed Riemann surfaces. Any naive set of Feynman rules led to multiple or infinite overcounting of surfaces. Given a Riemann surface one must to be able to associate to it a unique . This is achieved with a generalized minimal area problem [8] which asks for the (conformal) metric of least possible area under the condition that all nontrivial closed curves on the surface be longer than or equal to some fixed length, conventionally chosen to be 2π. Staring at the surface with its minimal area metric one can reconstruct the Feynman diagram from which it arises ( 6.4). The idea is that a minimal area metric must have closed geodesics of length § 2π that foliate the surface. These geodesics form foliation bands, that is, annuli foliated by homotopic geodesics. The height of a band is defined to be the shortest distance (on the band) between its boundaries. In contrast with metrics that arise from Jenkins-Strebel quadratic differentials [29], where the geodesics (horizontal trajectories) intersect in zero measure sets

4 † (critical graphs), the minimal area metrics can have bands of geodesics that cross. A band of height greater or equal to 2π cannot be crossed by any other band of geodesics and must be isometric to a flat cylinder of circumference 2π [15,16]. These are the propagators. The bands of infinite height are the external legs of the diagram. If the surface has no finite height foliation of height bigger than 2π, then the whole string diagram corresponds to an elementary interaction , where n 0 is the number of infinite height foliations, and g is the genus Vg,n ≥ of the surface. On each propagator, which is a flat cylinder, we mark two closed geodesics, each a distance π from each boundary of the cylinder. We also mark one curve, a distance π from the boundary on the semi-infinite cylinders. The string diagram is cut open along all these curves and the surface breaks up into a number of semi-infinite cylinders, the external legs, a number of finite cylinders, the propagators, and a number of surfaces with boundaries, that correspond to the elementary interactions. Those surfaces have ‘stubs’ of height π on each of their boundaries, arising from the height π foliation that was left attached from the bands that gave rise to the propagators. It is interesting that a foliation of height less than 2π, even though it may look like a propagator (if it happens not to be crossed by another foliation), must not be considered as such. It is as if, due to the minimum length condition of 2π on the string diagram, propagation by a distance smaller than 2π does not qualify as free propagation. The closed geodesics correspond to the closed strings, and the fact that the bands of foliations cross, for genus one and above, may be interpreted as saying that it is not clear where the string is. This crossing is necessary for modular invariance, and only happens within the elementary vertices. The difficulty with modular invariance in closed string theory was due to our inability to tell in which direction the strings propagate. This ambiguity does not exist for the minimal area string diagrams; propagators are unambiguously determined.

The uses of Stubs and the BV Equation. The role of the stubs is to prevent the appearance of short curves upon sewing. Since we leave a stub of length π on every boundary, upon gluing two boundaries and creating new closed curves, these will be longer than or equal to 2π. The stubs allowed us to characterize the elementary interactions of the complete quantum theory of closed strings as the surfaces with no finite height foliation of length greater than 2π. The stubs may not be essential for the classical theory since sewing in tree configurations does not introduce short curves. That is why, at the classical level only, we can specify the elementary interactions as the surfaces with no finite height foliations; these are the restricted polyhedra of [1,2]. Such interactions, which are contact-type overlaps, would be expected to be background independent [30] in the sense discussed recently in [31]. Then, as in open string field theory, the only background dependent ingredient of the classical theory would be the BRST operator. On the other hand stubs may be relevant in the classical theory, for using stubs string products are well defined for a set of string fields larger than that for which overlap- type string products exist. Furthermore, stubs are helpful in closed superstring field theory to prevent the collission of picture changing operators [32]. At the quantum level the stubs are essential and have the great virtue of making the full quantum BV equation true without

One such example was given in Ref.[16], Fig.5. †

5 need of regularization (in usual formulations of particle field theory, or in string theory with contact interactions, the ‘delta’ term in the BV equation needs regularization.) The closed string field theory master action amounts to a perturbative solution of the BV equation. It is of great interest to understand if this series solution has a finite radius of convergence or it is an asymptotic expansion. The case for the latter possibility has been emphasized by Shenker, who also indicated the possible relation with nonperturbative effects in string theory [33].

On Vacuum graphs. The BV master equation does not give us any constraint on the vacuum graphs of the string field theory. This is so because in the BV equation the master action always appears differentiated with respect to fields (or antifields). The master equation only guarantees the gauge independence of the S-matrix, and therefore it does not ensure that the string field theory will give the values of vacuum graphs that one would expect. Our geometrical recursion relations ( 5 ) determine the ‘vacuum vertices’ of the string field theory for genus § greater or equal to two. In this paper we will include in the master action theh ¯ dependent constants that arise from the vacuum vertices. This is done, even though the constants drop out of the master equation, because it seems likely that such constants would actually arise from a fundamental underlying action (once we find it!) upon expansion around a classical solution. The case of genus one is quite puzzling. The geometrical recursion relations do not determine this vacuum vertex. If we would compute this vacuum graph using just the propagator, we would get an infinite cover of the moduli space. If we insist on getting the genus one vacuum graph to work we must use a cutoff propagator instead of stubs on the vertices. The cutoff-propagator would include tubes with length in the interval [2π, ] rather ∞ than in the interval [0, ]. If we do this there is no need for stubs; the price we pay is a ∞ somewhat less natural kinetic term, which would contain, in addition to the BRST operator, + an insertion of exp(2πL0 ). It seems clear that we need further understanding of background independence in order to clarify the issues above.

Let us now indicate briefly the contents of this work. Sections 2, 3 and 4 are essentially § § § algebraic. In 2 we discuss all the conventions that we will use and derive basic relations § necessary for the later sections. We give a detailed discussion of Hermitian conjugation, BPZ conjugation [34], and inner products. We discuss sewing kets and bras (the so-called reflector states), their BRST properties and their exchange symmetry properties. We elaborate on the properties of basis states and their conjugate states. In 3 we study the string field and the § kinetic term. We explain in detail the subsidiary conditions on the string field, the ghosts number conditions, and Grassmanality. We discuss target space ghost number. The kinetic term is given and its basic properties are established. We also specify bases for fields and antifields, give the conventions for the BV formalism and verify that the kinetic term satisfies the master equation. In 4 the subject is the algebraic structure of the string field theory. We § state the properties of the string products and give the main identity satisfied by the products, along with examples. We also introduce the multilinear functions and give their properties. We then give the classical action, the field equation, prove gauge invariance, and compute the

6 ⋆ commutator of two gauge transformations. Finally we give the quantum master action, show it satisfies the quantum master equation, derive the master transformations and the BRST transformations. This section concludes with a mathematical definition for the L∞ algebra and an analysis of the algebraic structure that would arise if closed string field theory were constructed around a non-conformal background. Sections 5 and 6 deal with the geometrical issues that underlie the formulation of the § § string field theory. In 5 we define precisely the notion of string vertices (which are sets § Vg,n of Riemann surfaces), and state the basic conditions they must satisfy. We then discuss and elaborate on the fundamental geometrical relation [7], satisfied if the string vertices, via the Feynman rules, generate a decomposition of the relevant moduli spaces. In 6 we show how § the minimal area problem determines the subsets . Our discussion is based fully on the Vg,n properties of the geodesics in minimal area metrics. We state and prove a cutting theorem, which together with the results of [16] completes the study of the sewing properties of minimal area metrics. We also state what remains to be proven in order to have complete mathematical treatment of minimal area metrics. Section 7 combines the geometrical and the algebraic structures in order to construct § the string products and multilinear functions and prove their basic properties. This requires a detailed analysis of Schiffer variations (deformations of punctured Riemann surfaces with local coordinates) and the construction of a set of differential forms on the space (the Pg,n space of surfaces of genus g and n punctures, with local coordinates at the punctures specified b up to phases). Using heavily results of the operator formalism of Alvarez-Gaume, et.al. [36], further developed and clarified by Nelson and others [37], we show that the differential forms are well defined if the off-shell states that label the forms satisfy the subsidiary conditions − − b0 = L0 = 0. The BRST operator is shown to act as a total derivative on the forms defined in . We define the string products and the string multilinear functions and establish some Pg,n of their simple properties. Finally we prove the hermiticity of the string field vertices. b In section 8 we establish the main identity satisfied by the string field products, this § involves a detailed discussion of sewing of string field vertices. Section 9 is the final section. § It shows how to obtain the algebraic structures on physical states from the algebraic structures on the whole state space.

The present paper is long because of the large amount of material covered and because I wished to present a complete and useful set of conventions for closed string field theory along with detailed proofs of the most important properties. The material presented here is about 40% completely new; about 40% is proofs of results only sketched or announced before, and 20% is background material included for completeness.

⋆ This commutator has been independently obtained by Schubert [35].

7 2. Conformal Field Theory, Conjugation and Inner Products In this section we will give the conventions with which we will formulate the string field theory and some of the basic ingredients of the formalism. We begin with a description of the relevant ideas of conformal field theory, and a study of the ghost sector and its vacuum structure. Then we turn to defining hermitian conjugation and the associated antilinear inner product. Next we consider BPZ conjugation and introduce the associated linear inner product. These are essential to be able to write actions. We discuss bases of states for conformal field theory and derive a series of basic properties that will be essential to discuss the string field theory. The first work emphasizing the use of conformal field theory to define string field theory is that of LeClair, Peskin and Preitschopf [38].

2.1. Conformal Field Theory

Bosonic string field theory is formulated using a conformal field theory. This conformal field theory has a matter sector, which is some conformal field theory of total central charge equal to (+26) and a ghost sector, which is the conformal field theory of the reparametrization ghosts and has central charge equal to ( 26). The total conformal field theory has therefore − total central charge equal to zero. Consider the conformal field theory formulated in the z-plane with z = exp(τ + iσ). While the whole construction of the string field theory can be carried out abstractly, we can describe very concretely the ghost sector of the conformal field theory. Consider the case when the conformal field theory has holomorphic and antiholomorphic fields φ(z), φ(z) and stress tensors T (z), T (z). The field φ(z) is said to be a primary field of conformal dimension d if d 1 T (z)φ(w) φ(w)+ ∂ φ(w)+ , (2.1) ∼ (z w)2 z w w ··· − − corresponding to the transformation law

φ′(z′)(dz′)d = φ(z)(dz)d. (2.2)

The field is expanded in a basis of operators (oscillators) in the form φ dz φ(z)= n , φ = zn+d−1φ(z). (2.3) zn+d n 2πi Xn Z The operator φ is of dimension ( n) (recall z counts as dimension minus one). The total n − stress tensor, has central charge zero and is a dimension two field (it transforms as a quadratic differential on a surface). Its oscillator expansion is given by

L L T (z)= n , T (z)= n . (2.4) zn+2 zn+2 Xn Xn

The SL(2,C) vacuum, denoted by 1 , is the vacuum in the asymptotic past and it is | i associated with z = 0, with no operator inserted there. It follows from (2.3) and regularity

8 that φ 1 =0 for n d +1. (2.5) n| i ≥ − One introduces a state in the dual space, the SL(2,C) vacuum in the asymptotic future z = , ∞ denoted by 1 . It also follows from (2.3) (and (2.2) for z′ =1/z) that h |

1 φ =0 for n d 1. (2.6) h | n ≤ −

The canonical example of matter field theory is the theory of 26 free bosons i∂Xµ(z), µ i∂X (z), of dimensions (1, 0) and (0, 1) respectively (the first entry in ( , ) refers to the di- · · mension with respect to the holomorphic part of the stress tensor, and the second gives the dimension with respect to the antiholomorphic part of the stress tensor). The mode expansions are: µ µ α µ α i∂Xµ(z)= n , i∂ X (z)= n . (2.7) zn+1 zn+1 Xn Xn

Here the oscillators satisfy the algebra

µ ν µ ν µν [αn, αn] = [αn, αn]= ng δn+m,0, (2.8) where gµν is the Minkowskian metric ( , + +) (the case of compactified coordinates involves − ··· µ ν µ a few more subtleties, see for example the recent treatment in [31]). Moreover α0 = α0 = p . We then have b α 1 = α 1 =0 for n 0, and 1 α = 1 α =0 for n 0. (2.9) n| i n| i ≥ h | n h | n ≤

One actually defines a family of vacua 1,p , and dual vacua 1,p labelled by the eigenvalue | i h | of the momentum operator p via

b p 1,p = p 1,p , and 1,p p = 1,p p. (2.10) | i | i h | h | b b For these vacua the oscillators that annihilate it are given by

α 1,p = α 1,p =0 for n> 0, and 1,p α = 1,p α =0 for n< 0, (2.11) n| i n| i h | n h | n as follows from the two previous equations. Let us now consider the ghost conformal field theory. We have ghost fields c(z) and c(z) of dimensions ( 1, 0) and (0, 1) respectively, and antighost fields b(z) and b(z) of dimensions − −

9 (2, 0) and (0, 2) respectively. We therefore have c c¯ c(z)= n , c(z)= n , (2.12) zn−1 z¯n−1 Xn Xn

b ¯b b(z)= n , b(z)= n . (2.13) zn+2 z¯n+2 Xn Xn The modes satisfy the anticommutation relations b ,c = ¯b , c¯ = δ , (2.14) { n m} { n m} m+n,0 with all other anticommutators equal to zero. It is convenient to define new zero modes by linear combinations of the old ones 1 c± = (c c¯ ), b± = b ¯b . (2.15) 0 2 0 ± 0 0 0 ± 0 It follows from equations (2.5) and (2.6) that

bn 1,p = ¯bn 1,p =0 for n 1 | i | i ≥ − (2.16) c 1,p =¯c 1,p =0, for n 2, n| i n| i ≥ and 1,p bn = 1,p ¯bn =0 for n 1, h | h | ≤ (2.17) 1,p c = 1,p c¯ =0, for n 2. h | n h | n ≤ − We define the first quantized ghost number operator G by ∞ 1 G =3+ (c b b c )+ (c b b c ) + a.h. , (2.18) 2 0 0 − 0 0 −n n − −n n  nX=1 where a.h. denotes an identical contribution from the antiholomorphic sector. We included the constant shift of (+3) in order to have G 1,p = 0, that is, the ghost number of the SL(2,C) | i vacuum is zero. Due to the presence of two zero modes the state space space breaks into four sectors built on top of four different vacua forming a representation of the algebra of zero modes. These are defined as follows: ,p c c¯ 1,p , |↓↓ i ≡ 1 1| i ,p c+ ,p , |↑↓ i ≡ 0 |↓↓ i (2.19) ,p c− ,p , |↓↑ i ≡ 0 |↓↓ i ,p c+c− ,p . |↑↑ i ≡ 0 0 |↓↓ i The state ,p is a state of ghost number (+2) and it is annihilated by b±. This state is |↓↓ i 0 annihilated by all b and c with n 1 (similarly for the antiholomorphic oscillators). From n n ≥ left to right, the first entry in the kets refers to the (+) zero modes and the second entry refers to the ( ) zero modes. −

10 The conformal field theory stress tensor is the sum of the matter stress tensor (Tm(z), T m(z)), and the ghost stress tensor (Tg(z), T g(z)). The latter is given by T (z)= 2b(z) ∂c(z) ∂b(z) c(z), (2.20) g − · − · and the basic operator product expansion is 1 b(z)c(w) . (2.21) ∼ z w − The BRST operator of the theory is given by dz 1 dz 1 Q = c(z) Tm(z)+ Tg(z) + c(z) T m(z)+ T g(z) , (2.22) Z 2πi 2 Z 2πi 2   where the operators in the integrand are normal ordered. Using the above three equations, one can verify that Q, b(z) = T (z)+ T (z)= T (z), Q, b(z) = T (z)+ T (z)= T (z). (2.23) { } m g { } m g This equation implies that Q is of the form Q = c L + c L + , (2.24) 0 0 0 0 ··· where the dots indicate terms that do not involve the zero modes of the ghost fields. If we define L± = L L , (2.25) 0 0 ± 0 this together with (2.15) imply that (2.24) can be written as Q = c+L+ + c−L− + . (2.26) 0 0 0 0 ···

The contribution of the ghosts to L0 is given by ∞ L = 1+ n(b c + c b ), (2.27) 0(g) − −n n −n n Xn=1 with a similar expression for the antiholomorphic part. We then have that L0(g) and L0(g) annihilate the vacua L 1,p = L 1,p =0. (2.28) 0(g)| i 0(g)| i The ,p vacua are tachyonic as far as the ghost energy momentum is concerned, one has |↓↓ i L ,p = L ,p = 1 ,p . (2.29) 0(g)|↓↓ i 0(g)|↓↓ i − · |↓↓ i Since, for the usual bosonic string theory L + L = p2 + , where the dots correspond 0(m) 0(m) ··· to terms that annihilate the ,p vacua, one has that the total L+ operator on these vacua |↓↓ i 0 gives L+ ,p =(p2 2) ,p . (2.30) 0 |↓↓ i − |↓↓ i

11 2.2. Hermitian Conjugation, BPZ conjugation and Inner Products

Given a conformal field theory we can define an antilinear inner product, arising from Hermitian conjugation and a bilinear inner product, arising from BPZ conjugation. Both types of conjugations relate states built on the in-vacuum to states built on the out-vacuum. In order to be able to define these products we need to define a basic overlap. For the conformal field theory giving rise to bosonic string theory, we take 1,p c c¯ c+c−c c¯ 1,p′ = (2π)d δd(p p′), (2.31) h | −1 −1 0 0 1 1| i · − where δd is the d-dimensional delta function. The absence of an i in the right hand side, while very convenient to avoid unnecessary factors of i, will result in a slightly unusual formula for hermitian conjugation. For other conformal field theories one may take different definitions for the basic overlap. Hermitian conjugation and ( , ). Hermitian conjugation will be relevant to us since there is a · · need to impose a reality condition on the string field, as we will see in 3.1. Let us consider § for concreteness the bosonic string. The zero mode operators for the matter, namely the x’s and p’s, are both hermitian operators: x† = x and p† = p. One also has b ipx † −ipx b 1,p = e 1 , andb b1,p =(b 1,pb ) = 1 e , (2.32) | i | i h | | i h | which, in agreement with our earlierb definition, implies 1,p p = 1,pbp. One defines for all h | h | the oscillators (all n) † † † b αn = α−n, cn = c−n, bn = b−n, (2.33) and similarly for the antiholomorphic oscillators. Finally the hermitian conjugate of a general state is defined via † A A 1,p c c 1,p A† A†, (2.34)  1 ··· n| i  ≡ h | n ··· 1 where the A’s represent oscillators, c is a number, and the bar denotes complex conjugation. We will denote the hermitian conjugate state corresponding to A as A . The antilinear | i h hc| inner product is defined as follows: given two ket states A and B one defines | i | i (A, B) A B . Antilinear Inner Product (2.35) ≡h hc| i The label ‘hc’ is supressed for momentum eigenstrates; p will always denote the hermitian h | conjugate of p . As desired, the inner product is antilinear in the first argument and linear in | i the second ∗ (Aα + A′α′, B)= α∗(A, B)+ α′ (A′, B), (2.36) (A, Bβ + B′β′)=(A, B)β +(A, B′)β′. It follows from the definition of the basic overlap in (2.31) that (A, B)† = (B, A), (2.37) − or in other words ( A B )† = B A . The unusual minus sign arises due to the absence h hc| i −h hc| i of an i factor in the right hand side of (2.31), and because the hermitian conjugate of the

12 + − (c−1c−1c0 c0 c1c1) is minus itself. We have used the dagger in the left hand side of (2.37) instead of complex conjugation in order to take into account the fact that the c-numbers that multiply the states may have nontrivial statistics. Hermitian conjugation of an operator is defined in analogy to (2.34) and clearly gives a normal ordered operator if we start with one. We then have, by construction, that ( A, B)=(A, †B). (2.38) O O The BRST operator can be shown to be hermitian, and we therefore have (QA, B)=(A, QB), (2.39) for any two states A, B (of whatever statistics). Note finally that we can use the inner product to define a norm in the usual way. Equation (2.31) implies that only states of ghost number three have nonvanishing norm. The inner product we have defined pairs states of ghost numbers that add up to six. In terms of the first ⋆ quantized ghost number operator G introduced earlier, the ghost number of 1 is six, any h | ghost oscillator acting on it subtracts one unit of ghost number and any antighost oscillator adds one unit. It is actually more convenient to think of the ghost number of 1 as zero, then h | any ghost (antighost) oscillator adds (subtracts) one unit of ghost number, just as for states. In this convention, that we will follow throughout, the ghost number of a state does not change under hermitian conjugation. BPZ conjugation and , . There is another type of conjugation which will be most useful in h· ·i our writing of actions. It is BPZ conjugation [34], and relates in-states to out-states. Given a state A = A(0) 1 , with A a normal ordered operator, one defines the associated BPZ | i | i conjugate state A 1 I A(0), (2.40) h |≡h | ◦ † where I denotes the conformal mapping I(z)=1/z . The conformal mapping acting on the operator is defined as I A(z, z) A′(z′(z), z′(z)) where z′ = I(z), z′ = I(z) and A′(z′, z′) is ◦ ≡ related to A(z, z) via the conformal transformation properties of the operator. For the mode expansion (2.3) one takes ′ ′ T dz ′n+d−1 ′ ′ dz ′n+d−1 1 1 d φ I φn = z φ (z )= z − φ =( 1) φ−n, (2.41) n ≡ ◦ Z 2πi Z 2πi ′2 z′ − z   where we have used the superscript (T ) to denote this type of conjugation. The above formula enables us to calculate the BPZ conjugate of a string of oscillators: (c a b c 1 )T = 1 c (a )T (b )T (c )T (2.42) · n m p ···| i h | · n m p ··· Note that the order of the operators does not change and that the c-number is not complex conjugated. The ghost number of the BPZ conjugate state is the same as that of the original

⋆ Note that G is of the form G =3+ , where is an antihermitian operator. The choice of 1/z instead of the standardG oneG of 1/z in BPZ is more convenient for closed strings. † −

13 state. It follows from (2.41) that pT = p, and therefore we must take ( 1,p )T = 1, p The − | i h − | reader can check that the BPZ conjugates of the various ghost vacua ((2.19)) are given by b b ,p 1, p c−1c¯−1, h↓↓ |≡h − | + ,p 1, p c c−1c¯−1, h↑↓ |≡−h − | 0 (2.43) ,p 1, p c−c c¯ , h↓↑ |≡−h − | 0 −1 −1 ,p 1, p c+c−c c¯ . h↑↑ |≡h − | 0 0 −1 −1 Using BPZ conjugation one defines an inner product which is linear on both arguments

A, B A c− B , Linear Inner Product. (2.44) h i≡h | 0 | i − The utility of including the factor of c0 in the inner product arises because our string fields will − be defined to be annihilated by b0 , and the naive inner product of such states would vanish. In order to elucidate the properties of this inner product let us discuss the so-called conjugate states. Given a complete basis of states in a conformal field theory Φ , where r is a label, which {| ri} may include indices and/or continuous parameters, one defines the conjugate states Φc (not {h r|} hermitian conjugates nor BPZ conjugates) via the condition

Φc Φ = δ . (2.45) h r| si rs The conjugate states form a basis for all the states built on the out-vacuum (bra-states). Let us understand more explicitly this conjugation. We can construct a basis by using states built with oscillators. For example, the state conjugate to the vacuum is

1c = (2π)−d 1 c c c+c−c c , (2.46) h | h | −1 −1 0 0 1 1 It then follows by BPZ conjugation that 1c = (2π)−dc c c+c−c c 1 and that 1 1c = 1. | i −1 −1 0 0 1 1| i h | i Doing the same operations with an arbitrary state, we have

Φ =(c c )(b b )(a a ) 1 , | si −n1 ··· −ni −m1 ··· −mj −p1 ··· −pk | i c 1c Φs = (apk ap1 )(cmj cm1 )(bni bn1 ), h | h | ··· ··· ··· (2.47) Φ =( )i+k 1 (c c )(b b )(a a ), h s| − h | n1 ··· ni m1 ··· mj p1 ··· pk Φc =( )k+j(a a )(c c )(b b ) 1c . | si − −pk ··· −p1 −mj ··· −m1 −ni ··· −n1 | i c (i+j) It follows from the above that Φs Φs =( 1) , but (i + j) simply denotes the statistics of h | i − ‡ the state Φ . We therefore infer that corresponding to (2.45) we have | si Φ Φc =( )Φr δ . (2.48) h r| si − rs Whenever a field appears in the exponent of a sign factor, as in ( )Φ, it denotes what is usually called ‡ the Grassmanality ǫ(Φ) of the field Φ. This Grassmanality ǫ(Φ)− is an even integer, if Φ is Grassmann even, or an odd integer, if Φ is Grassmann odd.

14 Completeness of the states implies that the identity operator can be written as

1 = Φ Φc = Φc Φ ( )Φr , (2.49) | rih r| | rih r| − Xr Xr as can be checked by multiplying arbitrary states from the left and from the right, and making use of (2.45) and (2.48). We claim now that for arbitrary states A and B of definite statistics | i | i one has A B = B A ( 1)AB. (2.50) h | i h | i − This is established by writing A = a Φc and B = Φ b , where a and b are h | sh s| | i s | si s s s c-numbers, possibly anticommuting. ItP then follows that A PB = asbs, and a short com- (as+Φs)(bs+Φs) h | i AB putation shows that B A = asbs( ) where the signP factor is simply ( ) , h | i − − since the states were assumedP to be of definite statistics.

The Reflector States. We can use the conjugate states to give a simple expression for a state that implements BPZ conjugation. One defines

R = Φ Φc , (2.51) h 12| 2h r| 1h r| Xr where following our notation Φ denotes the BPZ conjugate of Φ , and the subscripts on h r| | ri the bras are the labels for the Hilbert spaces. The bra R may be thought as an object in the h | tensor product of two Hilbert spaces (it defines a multilinear function on a pair of states by the operation of contraction). It was constructed so that

R A = A , and R A = A , (2.52) h 12| i1 2h | h 12| i2 1h | the first of which follows directly from expanding the state A along the basis Φ and using | si (2.45), and the second by expanding A along the basis Φc and using (2.48). This suggests | si that R is actually symmetric in the labels 1 and 2: h 12| R = R . (2.53) h 12| h 21| Rather than prove this, it is useful to derive slightly more general rearrangement properties:

Φ Φc = Φ Φc , | si1| si2 | si2| si1 G(ΦXs)=G G(ΦXs)=6−G (2.54) Φ Φc = Φ Φc . 1h s| 2h s| 2h s| 1h s| G(ΦXs)=G G(ΦXs)=6−G

To establish the first relation we begin with the left hand side and introduce two complete

15 sums over states ( )Φp Φc Φ Φ Φ Φc Φc , − | pi1 h p| si| qi2 h q| si G(ΦXs)=G Xp,q (2.55) Φp ΦpΦq c c c = ( ) ( ) Φq 2 Φp 1 Φp Φs Φq Φs , G G − − | i | i h | ih | i G(ΦXs)=G (ΦXp)=6− G(Φq )=6−G where we used the fact that the overlap only couples states whose ghost numbers add up to six to restrict the sums to a space of fixed ghost number, and we moved the state Φ all | qi the way to the left. Note that since Φ Φ is an ordinary number, the statistics of Φ is the h p| si p same as that of Φs for all nonvanishing terms in the above sum. We can therefore replace ( )ΦpΦq Φc Φc by ( )ΦsΦq Φc Φc which equals Φc Φc , to find − h q| si − h q| si h s| qi ( )Φp Φ Φc Φ Φ Φc Φc , (2.56) − | qi2| pi1 h p| sih s| qi G(ΦXp)=6−G Xs G(Φq )=6−G where we can now freely sum over all possible Φs since in any case the ones with ghost number different from G would vanish by ghost number conservation. Since the sum over Φs is simply the identity operator we find that the above reduces to

Φ ( )Φp Φc Φ Φc = Φ Φc , (2.57) | qi2 − | pi1 h p| qi | qi2| qi1 G(ΦXq )=6−G Xp G(ΦXq )=6−G where again we made the sum over states Φp unconstrained, since terms with other ghost numbers vanish, and then recognized the identity operator to obtain the result we were after. The second identity in (2.54) is established in an exactly analogous way. The symmetry R = R of the reflector is an immediate consequence of this second identity. h 12| h 21| We also have the useful relations (1) (2) R (cn + c )=0, h 12| −n (1) (2) R (bn b )=0, (2.58) h 12| − −n R (Q + Q )=0. h 12| 1 2 The first two identities, which hold for all n, can be obtained as special cases of (2.52). The last equation can be checked directly in particular cases, but follows from basic operator formalism ideas [36]; the reflector simply represents a two punctured sphere and contour deformation implies that the two BRST operators cancel each other. We can use the reflector to write the basic overlap as A B = R A B , (2.59) h | i h 12| i1| i2 and the inner product we defined before as

A, B = R A c−(2) B . (2.60) h i h 12| i1 0 | i2 Note that the overlap is nondegenerate, that is B A = 0 for all B implies A = 0. Further- h | i | i more, the inner product is nondegenerate too. First note that in A, B the terms in A and B h i

16 − − proportional to c0 drop out, and therefore A and B can be assumed to be annihilated by b0 . The statement of nondegeneracy is therefore

If B, A =0, B, and b− A =0, then A =0, (2.61) h i ∀ 0 | i | i

since the nondegeneracy of the overlap B, A = B c− A = 0 for all B implies c− A = 0, h i h | 0 | i 0 | i but then multiplying by b− and using b− A = 0 one readily sees that A = 0. The inner 0 0 | i | i product satisfies the two following properties

A, B =( )(A+1)(B+1) B, A , (2.62) h i − h i

− − and, for A and B annihilated by b0 and L0 , we have

QA, B =( )A A, QB . (2.63) h i − h i

Equation (2.62) follows directly from (2.60), the symmetry of R12 and the first relation in − (2.58) (used for c0 ). Equation (2.63) is less straightforward. First note that the left hand side can be written as ( )A+1 R A Q(2)c−(2) B , (2.64) − h 12| i1 0 | i2 using (2.58). We now introduce the factor 1 = b−(2),c−(2) into the above to find { 0 0 }

( )A+1 R A (b−(2)c−(2) + c−(2)b−(2)) Q(2)c−(2) B . (2.65) − h 12| i1 0 0 0 0 0 | i2

−(2) The first term in the parenthesis vanishes because b0 can be pushed all the way to the left until it hits the reflector, becomes b−(1) and kills A . The second term gives 0 | i1

( )A+1 R A c−(2) b−(2), Q(2) c−(2) c−(2)Q(2) b−(2),c−(2) B , (2.66) − h 12| i1  0 { 0 } 0 − 0 { 0 0 }| i2 where we used b− B = 0. The first anticommutator gives L−, which commutes with c− and 0 | i 0 0 annihilates B , therefore the only remaining term is | i

( )A R A c−(2)Q(2) B =( )A A, QB , (2.67) − h 12| i1 0 | i2 − h i as we wanted to show.

17 Related to the reflector R , we introduce a ket reflector R defined by h | | i R Φ Φc = Φ Φc , (2.68) | 12i ≡ | si1| si2 | si2| si1 Xs Xs where the equality, implying R = R , is an immediate consequence of the first equation | 12i | 21i in (2.54). It follows that

R R = Φ Φc Φ Φc = Φ Φc = ( )Φs Φc Φ , (2.69) h ij| jki ih r|jh r| sij| sik ih s|·| sik − | sik · ih s| Xr,s Xs Xs where the last expression to the right is simply the identity operator (see Eqn. (2.49)) except for the fact that the labels of the Hilbert spaces do not coincide. This operator simply relabels states (or equivalently, it maps to ) Hi Hk R R A = A . (2.70) h ij| jki| ii | ik In analogy with the properties of the reflector, this conjugate reflector satisfies (1) (2) (cn + c ) R =0, −n | 12i (1) (2) (bn b ) R =0, (2.71) − −n | 12i (Q + Q ) R =0. 1 2 | 12i It is useful for our future applications to introduce some further notation. We define the ‘tilde’ states which are simply related to the conjugate states by multiplication by the antighost − zero mode b0 Φ b− Φc . (2.72) | si ≡ 0 | si We also introduce the tilde reflector e R b−(1) R = b−(2) R , (2.73) | 12i ≡ 0 | 12i 0 | 12i where the equality follows frome (2.71). This reflector can therefore be written as

R = b−(2) Φ Φc = ( )Φs Φ Φ . (2.74) | 12i 0 | si1| si2 − | si1| si2 Xs Xs e e It should be noted that this reflector does not satisfy a simple BRST property; we have (Q + Q ) R = 0. It is useful to introduce the ‘primed’ reflectors, R′ and R′ which 1 2 | 12i 6 | 12i | 12i are simply the reflectors but with the sum over states restricted to states annihilated by L− e e 0 R′ = R , R′ = R , (2.75) | 12i P1P2| 12i | 12i P1P2| 12i where , given by e e Pi (i) dθ iθ(L(i)−L ) i = e 0 0 , (2.76) P Z 2π is the projector, in the i-th , to states annihilated by (L L ). It satisfies 0 − 0 L− = L− = 0. Since the BRST operator satisfies Q, b− = L−, and commutes with , P 0 0 P { 0 } 0 P

18 we have that Q, b− = 0. As a consequence the primed reflectors satisfy simple BRST { 0 P} properties (Q + Q ) R′ =0, (Q + Q ) R′ =0. (2.77) 1 2 | 12i 1 2 | 12i We can write the primed reflectors as e

′ ′ R′ = Φ Φc , R′ = ( )Φs Φ Φ , (2.78) | 12i | si1| si2 | 12i − | si1| si2 Xs Xs f e − where the primed sums remind us that we are only summing over states annihilated by L0 . − − Note that since the conjugate of a state annihilated by L0 is also annihilated by L0 , it is actually enough to have one projector rather than two in each of the expressions in (2.75). Thus we write R′ = R , R′ = b−(1) R . (2.79) | 12i P1| 12i | 12i 0 P1| 12i e We give, for illustration, the oscillator form of the reflector state for the bosonic string:

dp +(1) +(2) −(1) −(2) R12 = 1 ,p 2 , p exp(E12)(c + c )(c + c ), (2.80) h | Z (2π)d h↓↓ | h↓↓ − | 0 0 0 0

where 1 (1) (2) (1) (2) (1) (2) E = αn αn + cn bn bn cn + a.h., (2.81) 12 n · −  Xn≥1 with a.h. denoting the antiholomorphic part. This state corresponds to a two punctured sphere with local coordinates z (z)= z around z = 0, and z (z)=1/z around z = . 1 2 ∞ A Rearrangement Property. We conclude this section by establishing a result that will be needed in the following sections. We first show that

− − c −1 Φ ′ Φ = αc Φ ′ , with b Φ ′ =0, implies Φ =(α) ( ) s Φ ′ . (2.82) | si 0 | s i 0 | s i | si − | s i c e c This is proven by first realizing that under the assumptions Φs must be of the form Φs = − c | i −1 Φs′ c | i b χ . Then one requires that Φ Φ = δ to find that χ = (α) ( ) Φ ′ . The 0 | i h s| ri rs | i − | s i rearrangement property that we wish to establish reads

f(G) Φ Φ = f(G) Φ Φ . (2.83) | si1O| si2 − | si1O| si2 G(ΦXs)=G G(ΦXs)=5−G e e Here stands for some arbitrary operator, and f(G) is some arbitrary function of the ghost O number. By construction, the sum of states on the left hand side runs over all states Φs of the − specified ghost number, and annihilated by b0 (since otherwise they would be annihilated).

19 − The equation also holds if the sum is restricted to states annihilated by L0 . The left hand side of this equation can be written as

= f(G)( )(O+1)Φs b−(2) Φ Φc = f(G)( )(O+1)Φs b−(2) Φ Φc , − O 0 | si1| si2 − O 0 | si2| si1 G(ΦXs)=G G(ΦXs)=6−G (2.84) −(2) where use was made of (2.54). Notice now that Φ must be of the form c Φ ′ and | si2 0 | s i2 therefore we can use (2.82) to replace

−(2) c −(2) −(2) Φ ′ Φ ′ b Φ Φ = b c Φ ′ ( ) s Φ ′ = Φ ′ Φ ′ ( ) s , (2.85) 0 | si2| si1 0 0 | s i2 − | s i1 | s i2| s i1 − e e and therefore back in (2.84) we have

(O+1)(Φ ′ +1) Φ ′ = f(G)( ) s Φ ′ Φ ′ ( ) s , (2.86) − O| s i2| s i1 − G(Φ ′ )=6X−G−1 s e where we shifted the ghost number and wrote the sign factors in terms of the new set of states we are summing over. Since a state and its tilde conjugate commute we write

OΦ ′ = f(G)( ) s Φ ′ Φ ′ = f(G) Φ ′ Φ ′ , (2.87) − − O| s i1| s i2 − | s i1O| s i2 G(ΦX′ )=5−G e G(ΦX′ )=5−G s e s e which is the result we wanted to establish.

3. The String Field and Its Kinetic Term

In this section we begin our discussion of string field theory by describing the string field, and discussing in detail the subsidiary conditions it must satisfy, its reality condition, statistics, and ghost number. It is useful to distinguish between first quantized ghost number and target space ghost number. We show that the ghost number anomalies of the first quantized path integrals precisely imply that a suitably defined target space ghost number is conserved. We are then able to write the kinetic term for the closed string field theory and to show its hermiticity and its gauge invariance. The semirelative is seen to be the type of cohomology that describes the physical spectrum. Finally, we review the basics of the Batalin-Vilkovisky formalism, give the master equation, specify a basis for fields and antifields in closed string field theory and prove that the kinetic term satisfies the master equation.

20 3.1. The Dynamical String Field

⋆ An arbitrary string field is a vector in the Hilbert space of the conformal field theory. Given a basis of states Φ , an arbitrary string field is simply an arbitrary linear superposition of {| si} the basis states: Ψ = Φ ψ , (3.1) | i | si s Xs Here each target space field ψ is the component of the vector Ψ along the basis vector Φ . s | i | si The target space fields are in general complex numbers, and may be Grassmann even or odd. The dynamical string field is the dynamical variable of the string field theory, that is, the string field that enters into the action. This string field is not totally arbitrary. There are conditions limiting the possibilities in the general expansion above. These conditions are the subject of the present subsection.

− Subsidiary Conditions. We will demand that the closed string field be annihilated both by b0 − and L0 , namely (b b ) Ψ =(L L ) Ψ =0. (3.2) 0 − 0 | i 0 − 0 | i The necessity of such conditions is well understood by now; if we do not impose them we do not know how to write a kinetic term for the closed string field (see 3.2) nor how to write § interactions (see 7.3). The same conditions apply for the string field Λ which corresponds § − | i to the gauge parameter. Given the b0 condition, the closed string field can be expanded as

Ψ = φ , ψ + φ , ψ , (3.3) | i | s ↓↓i s | s ↑↓i s e where φ , denotes the subset of the basis states Φ which are built on the vacuum | s ↓↓i | si |↓↓i (and similarly for the second term).

Ghost Number G. The ghost number of a component of a string field is conventionally defined to be given by the ghost number of the (first quantized) state. Thus if we have G Φ = G Φ | si s| si we define G( Φ ψ ) (G Φ )ψ = G ( Φ ψ ). (3.4) | si s ≡ | si s s | si s For the classical theory the dynamical closed string field Ψ entering in the string action is | i taken to be of ghost number G = +2. For the general master action the string field Ψ | i contains fields of all integer ghost numbers.

Target Space Ghost Number gt. It is possible to assign to the target space fields a conserved ghost number gt. For a target space field ψ entering the string field as Φ ψ , we simply s | si s

⋆ Technically speaking, the state space of the matter plus ghosts conformal field theory form an indefinite metric Hilbert space, or simply an inner-product space.

21 define gt(ψ )=2 G . (3.5) s − s Thus in the classical action all spacetime fields have ghost number gt = 0, and in the master action we have target space fields of all ghost numbers. As a consequence of ghost number selection rules in the first quantized formalism, the interactions of target space fields will conserve target space ghost number. We can explain this now. When one defines the interaction of n string fields via Riemann surfaces of genus g one does it by use of a correlator on those surfaces. One must do an integral over a subset of the corresponding moduli space , Mg,n whose dimensionality is 6g 6+2n. Such dimensionality requires precisely the same number − of antighost insertions in the correlator. Moreover, in order to have a nonvanishing answer, the total ghost number inside the correlator must be 6 6g (= dim ). Thus we need that − − Mg the sum of ghost numbers Gi of the n string fields to be inserted be equal to 2n, namely, n G = 2n. This implies that gt = (2 G )=2n G = 0, which shows that i=1 i i − i − i Pthe ghost numbers of the target spaceP fieldsP add up to zero for everyP interaction in the master † action.

Grassmanality. The string field Ψ is declared to be grassman even. This will be true both | i for the classical action and for the master action. We also declare the vacua 1,p to be even | i (and similarly for the dual vacua). The grassmanality of the state Φ is defined to be the | si grassmanality of the conformal field theory operator Φs that creates the state by acting on the vacuum. This defines the grassmanality of the target space field ψs as that which together with the grassmanality of Φ makes up an even object. It is important to note that statistics | si and ghost number are correlated. If the ghost number G of a ket Φ is even, the ket must s | si be Grassmann even, and then the corresponding target space field has even target space ghost number and is Grassman even. Similarly if the ghost number G of a field Φ is odd, the ket s | si must be odd, and then the corresponding target space field has odd target space ghost number and is Grassmann odd. This implies that the interactions for target space fields in the master action have overall even statistics. In other words, the master action is a boson.

Reality Condition on the String Field. A priori our definition of the dynamical string field allows for a complex spacetime field associated to every state of the conformal field theory. This is clearly twice as much as we have in string theory, for example, the closed string tachyon is a real scalar field. It is convenient to impose the following reality condition, we demand that hermitian conjugation and BPZ conjugation give the same state up to a sign:

( Ψ )† Ψ = Ψ . (3.6) | i ≡h hc| −h | In formulating closed string field theory there is no necessity to investigate the implications of the reality conditions on the expansion of the string field in terms of component fields. In

This counting applies to all interactions (n 1). There is no interaction for a one punctured sphere. † For the two puncture sphere, which corresponds≥ to the kinetic term, the above requires ( 2) antighost − − insertions. As we will see, one inserts the operator c0 Q of ghost number two.

22 order to fix the normalization of the action in section 3.2, we will demand that the tachyon § field that appears conventionally in the string field as:

dp T = φ(p) c1c1 1,p , (3.7) | i Z (2π)d | i

appear in the closed string field theory action with standard normalization. The hermitian and BPZ conjugates of the tachyon are given by

dp ∗ Thc = φ (p) 1,p c−1c−1, h | Z (2π)d h | (3.8) dp T = φ(p) 1, p c−1c−1, h | Z (2π)d h − | and Eqn.(3.6) requires the standard reality condition φ∗(p)= φ( p). − Summarizing all the conditions on the dynamical string field we have

(b b ) Ψ =(L L ) Ψ =0, 0 − 0 | i 0 − 0 | i ( Ψ )† = Ψ , reality, | i −h | (3.9) Ψ is Grassmann even, | i G Ψ =2 Ψ , for the classical action. | i | i

3.2. The Kinetic Term

Having discussed in detail the relevant issues of conjugation and the conditions that must be imposed on the closed string field we now turn to the closed string field kinetic term

1 S = Ψ c−Q Ψ . (3.10) 0,2 2h | 0 | i

− Since the c0 Q operator has ghost number two, and the string field has ghost number two, the ⋆ total ghost number adds up to six, as it should. This shows that the BRST operator alone could not have been a suitable kinetic operator for any choice of ghost number for the string field. Thus the necessity of an extra insertion. In order to preserve gauge invariance in the − presence of the insertion of c0 the dynamical string field must satisfy the subsidiary conditions

⋆ Most recent works in closed string field theory use the convention where the string field Ψ is of ghost − | i number +3, and is annihilated by c0 . The convention chosen in the present work is useful to make explicit − factors of c0 dissapear from the identities satisfied by the string products.

23 b− Ψ = L− Ψ = 0. Using the reality condition on the string field and the definition of the 0 | i 0 | i antilinear inner product, the kinetic term can also be written as

1 S = (Ψ,c−QΨ), (3.11) 0,2 −2 0 and using the definition of the linear inner product, it can be written as

1 S = Ψ, QΨ , (3.12) 0,2 2h i which is the nicest expression for the kinetic term and justifies our introduction of the linear inner product with a ghost zero mode insertion.

Hermiticity of S0,2. We check this beginning with (3.11)

(Ψ,c−QΨ)† = (c−QΨ , Ψ), 0 − 0 (3.13) = (Ψ , Qc−Ψ), − 0 − − where we made use of (2.37) and of the hermiticity of c0 and Q. Since Q, c0 does not vanish, { − } we are not yet through. One must use the fact that Ψ is annihilated by b0 , and we do this by inserting the factor b−,c− = 1 in the above expression { 0 0 } − − − − − = (Ψ , [b0 c0 + c0 b0 ]Qc0 Ψ) − − − − = (Ψ , c0 b0 Qc0 Ψ) − − − − (3.14) = +(Ψ , c0 Q b0 ,c0 Ψ) − { } = +(Ψ , c0 QΨ)

− proving the hermiticity of the kinetic term. In the first step we used the hermiticity of b0 and − − b0 Ψ = 0. In the second step we moved the remaining b0 to the right until it annihilates the | i − string field. On the way one picks up the commutator with Q giving L0 which annihilates − − c0 Ψ, and the commutator with c0 , as shown in the equation. Field Equation. This time we begin with (3.12) and vary the string field to find

2 δS = δΨ, QΨ + Ψ, QδΨ , (3.15) · 0,2 h i h i using (2.62) the second term can be written as QδΨ, Ψ and using (2.63) we write it as h i δΨ, QΨ . Therefore the total variation is δS = δΨ, QΨ and the nondegeneracy of the h i 0,2 h i inner product (see (2.61)) together with the fact that b−Q Ψ 0 implies that the equation 0 | i ≡ of motion is simply Q Ψ =0. (3.16) | i This equation is simply what one would expect naively from (3.12).

24 Gauge Invariance. The gauge transformations of the classical theory are given by

δ Ψ = Q Λ , (3.17) | i | i where the gauge parameter must satisfy the following conditions

b− Λ = L− Λ =0, 0 | i 0 | i ( Λ )† = Λ , reality, | i −h | (3.18) Λ Grassmann odd, | i G Λ = Λ , ghost number, | i | i which follow from the corresponding conditions on the string field in Eqn. (3.9) (the second condition is verified by showing that QΛ = Λ Q and (QΛ) = Λ Q). The gauge param- h | h | h hc| h hc| eter can also be expanded in terms of target space gauge parameters as Λ = Φ λ . Since | i | ri r the ket Φr must have ghost number one, it must be Grassmann odd, and asP a consequence | i (Eqn.(3.18)) the target space gauge parameters are Grassmann even, as expected. The equation of motion in (3.16) is clearly gauge invariant under this gauge transformation. Let us check the gauge invariance of the action

2 δS0,2 = QΛ , QΨ + Ψ, QQΛ , · h i h i (3.19) = Λ , QQΨ)=0, −h

where use was made of Eqn. (2.63) and of the nilpotency of the BRST operator.

Physical Spectrum. As a consequence of the field equations and the gauge transformations the physical spectrum of the theory is defined by the semirelative cohomology

Q Ψ =0, Ψ Ψ + Q Λ , | i | i≡| i | i (3.20) b− Ψ = b− Λ =0. 0 | i 0 | i

− It is possible to show that all cohomology is always concentrated on the subspace L0 = 0, so this condition need not be imposed explicitly in the cohomology problem (it must be imposed explicitly on the string field in order to have a gauge invariant kinetic term). The cohomology computed with no subsidiary conditions on the string field nor gauge parameters is called absolute cohomology. The closed string relative cohomology is computed imposing the conditions that both b0 and b0 annihilate the string field and the gauge parameters.

Let us conclude by writing out the string field kinetic term for the case of the tachyon.

25 Using (3.7), (3.8), and (2.26) in the kinetic term (3.10) we find ′ tach 1 dp dp ′ ′ − + S = φ(p ) p , 1 c−1c−1c c (L0 + L0)c1c1 p, 1 φ(p), (3.21) kin 2 Z (2π)d Z (2π)d h− | 0 0 | i and using L + L = p2 2 (Eqn. (2.30)) and (2.31) we find 0 0 − 1 dp Stach = φ( p)(p2 2) φ(p), (3.22) kin −2 Z (2π)d − − which is indeed the correctly normalized tachyon kinetic term.

3.3. Batalin Vilkovisky Formalism and the Kinetic Term

In the Batalin Vilkovisky formalism one works with a set of fields ψs and a corresponding set ∗ of antifields ψs . The index s may possibly run over many indices, discrete and/or continuous. s s One encodes the grassmanality of ψ in ǫs; if ψ is Grassmann even (odd), ǫs is an even (odd) ∗ number. The grassmanality of ψs is similarly encoded in ǫs∗ . ∗ The fields and antifields are paired, with ψs denoting the antifield corresponding to the field ψs. The elements of a pair have opposite statistics and their ghost numbers add up to ( 1): − t s t ∗ ǫ ∗ = ǫ +1, g (ψ )+ g (ψ )= 1, (3.23) s s s − The master action is built using fields and antifields with the “boundary condition” that when the antifields are set to zero (andh ¯ is set to zero) we should recover the classical action. The full master action S must satisfy the complete Batalin-Vilkovisky master equation

∂rS ∂lS ∂r ∂lS s ∗ = h¯ s ∗ , (3.24) ∂ψ ∂ψs − ∂ψ ∂ψs where repeated indices are summed, and ∂r, ∂l denote derivatives from the right and from the ⋆ left respectively. If this equation is satisfied it follows that the S-matrix of the quantum theory is independent of the gauge fixing conditions [11]. This equation is sometimes written more briefly as 1 S,S = h¯∆S, (3.25) 2{ } − where the antibracket , and the ∆ operator are given by {· ·} ∂rG ∂lH ∂rG ∂lH ∂r ∂l G,H = s ∗ ∗ s , ∆= s ∗ . (3.26) { } ∂ψ ∂ψs − ∂ψs ∂ψ ∂ψ ∂ψs The antibracket satisfies a Jacobi identity of the form

( )(ǫF +1)(ǫH +1) F, G,H + cyclic = 0. (3.27) − { { }} We now define the “master transformations”, which are the precursors of the BRST trans-

s r r s r ǫ ǫ r ⋆ For η a c-number one has: ∂r/∂ψ (ηψ )= ηδ and ∂l/∂ψ (ηψ ) = ( ) s η ηδ . s − s

26 formations of the gauge fixed theory and of the gauge transformations. We denote the master trasformation of an object by s . They are defined by O O s = ,S µ, (3.28) O {O }· where is an arbitrary function of the fields and antifields, and µ is an anticommuting c- O number. It follows from this definition and that of the antibracket that

s ∂lS ∗ ∂lS sψ = ∗ µ, and sψs = µ. (3.29) ∂ψs · −∂ψs · Note that these transformations do not change the statistics of the object that is varied. Let us now perform two transformations in succession, one with parameter µ and the other with parameter µ′: s(s )= s ,S µ, O { O }· = ,S µ′,S µ, {{O }· }· (3.30) = ,S ,S µ′µ, {{O } }· , S,S µ′µ, ∼ {O { }}· where in the last step we used the Jacobi identity. If the action S satisfies the classical master equation S,S = 0 (theh ¯ = 0 reduction of (3.24)) then the master transformations are { } . Moreover, in this case, the action S itself is invariant under master transformations, since sS = S,S µ = 0. Thus: { }· S,S =0, implies s2 =0, and sS =0. (3.31) { } In the remainder of this subsection we will describe a basis for fields and antifields in closed string field theory, show that the kinetic term satisfies the classical master equation, and conclude by giving two selection rules showing what type of terms cannot appear in the master equation. The basic idea on how to decompose the string field into fields and antifields is due to Thorn who studied this problem for the case of open string fields [24]. What follows is simply an extension to closed strings. One decomposes the string field Ψ as | i Ψ = Ψ + Ψ , (3.32) | i | −i | +i where Ψ− contains all the fields and Ψ+ will contain all the antifields. Both Ψ− and Ψ+ | i − − | i | i | i must be annihilated by b0 and L0 . The fields appear as follows ′ Ψ = Φ ψs. (3.33) | −i | si G(ΦXs)≤2

The sum extends over the basis of states Φs with ghost numbers less than or equal to two, and | i − s the prime reminds us that the states are annihilated by L0 . The target space fields ψ have

27 target space ghost number gt(ψs)=2 G(Φ ). It follows that the fields include all possible − s values of gt 0. Half of the fields appear along states built on the vacuum and the other ≥ |↓↓i half along states built on the . Note that the complete set of fields is much larger than the |↑↓i set of fields in the classical action, we now include all possible positive ghost numbers. The antifields appear as ′ Ψ = Φ ψ∗, (3.34) | +i | si s G(ΦXs)≤2 e where ψ∗ is the antifield corresponding to the field ψs and Φ = b− Φc , as defined earlier. s | si 0 | si Since G(Φ )=5 G(Φ ), it follows that the ghost number of ψ∗ is s − s e s gt(eψ∗)=2 G(Φ )=2 (5 G(Φ )) = 3+(2 gt(ψs)) = 1 gt(ψs), (3.35) s − s − − s − − − − and therefore, the targete space ghost numbers of a field and its antifield add up to ( 1), as it − should be. Moreover, the statistics of the antifield is indeed opposite to that of the field, since the statistics of the corresponding first quantized states are opposite. A field corresponding to a state built on the vacuum is always paired with an antifield corresponding to a state |↓↓i built on the vacuum, and viceversa. Note that the ghost number of the antifields takes |↑↓i all possible negative values. It is rather interesting that the closed string master action (withh ¯ = 0), satisfying the classical master equation, is simply given by the same expression that gives the classical action, but without restriction on the ghost number of the string field. Indeed upon setting all the antifields to zero in such an expression one recovers the classical action. This happens because all fields have ghost number greater or equal to zero, and the only ghost number conserving terms are the ones involving the ghost zero fields, that is, the classical fields. Let us now verify that the above ansatz, for the case of the kinetic term, does indeed satisfy the master equation. The master kinetic term is given by 1 S = R Ψ Ψ , (3.36) 0,2 2h 12| 1iO2| 2i where, for brevity, we introduced the grassmann even operator given by = c−Q, which O O 0 satisfies R ( ) = 0, and now the string field contains all ghost numbers. It follows h 12| O1 − O2 by taking derivatives that 0 0 ∂rS2 ∂lS2 Φs+1 s = R12 Ψ1 2 Φs 2, and ∗ =( ) R34 Ψ3 4 Φs 4, (3.37) ∂ψ h | iO | i ∂ψs − h | iO | i e and therefore back into the classical master equation we have

0 0 ′ ∂rS2 ∂lS2 Φs s ∗ = R12 Ψ1 2 R34 Ψ3 4 Φs 2 Φs 4( ) , ∂ψ ∂ψs −h | iO h | iO | i | i − G(ΦXs)≤2 e (3.38) 1 ′ = R Ψ R Ψ ( )Φs Φ Φ + Φ Φ , −2h 12| 1iO2h 34| 3iO4 − | si2| si4 | si4| si2 G(ΦXs)≤2  e e

28 where we used the symmetry of the terms to the left of the sum under the exchange of the labels 2 and 4. We now use (2.83) to replace the second term in the sum

′ ′ ′ ( )Φs Φ Φ = ( )Φs+1 Φ Φ = ( )Φs Φ Φ , (3.39) − | si4| si2 − − | si4| si2 − | si2| si4 G(ΦXs)≤2 G(ΦXs)≥3 G(ΦXs)≥3 e e e and back into (3.38) we now obtain a sum extending over all states

1 ′ = R Ψ R Ψ ( )Φs Φ Φ , (3.40) −2 h 12| 1iO2h 34| 3iO4 − | si2| si4 Xs e and the sum is recognized to be simply a primed gluing ket (see Eqns. (2.78) and (2.79)). We therefore have 0 0 ∂rS2 ∂lS2 −(4) ′ 2 s ∗ = R12 Ψ1 2 R34 4 Ψ3 b0 R 24 , · ∂ψ ∂ψs −h | iO h |O | i | i −(3) = Ψ2 2 R34 b0 3 Ψ3 2 R24 , −h |O h | O | iP | i (3.41) = Ψ R b−(3) Ψ R , −h 2|O2h 34| 0 O3| 3i| 24i = Ψ b−(2) Ψ , −h 2|O2 0 O2| 2i = Ψ b− Ψ =0, −h |O 0 O| i where the projector 2 was moved all the way until it becomes one acting on Ψ2 , use was made P − − − − h | of Eqn.(2.70), and of the fact that b0 = c0 Qb0 c0 Q vanishes between states annihilated − O O by b0 . This concludes our verification that the kinetic term satisfies the master equation. Selection rules. It is clear from the master equation that field/antifield independent terms in the master action do not appear in the master equation. It seems possible, at first sight, that there are field independent terms in the master equation; for example, tadpole terms (interactions with one field) could give constants in the left hand side of the equation, and bilinear interactions could give rise to constants in the right hand side. We now show this is not possible, we claim that

g g g ∂rS1 ∂lS1 ∂r∂lS2 s ∗ =0, and s ∗ =0, (3.42) ∂ψ ∂ψs ∂ψ ∂ψs

g g where Sn (g 0) denotes the part of the master action S proportional toh ¯ and with n string ≥ fields. The reason these terms vanish is simple. When we have only one string field in the g interaction term (as in S1 ) only a target space field of ghost number zero can appear, thus a field. Then, the derivative with respect to the antifield must vanish. For the second case (that g of S2 ) one is dealing with terms with two string fields, the corresponding target space fields must therefore have opposite ghost numbers, thus they can never correspond to a field and its antifield.

29 4. Algebraic Structure of Closed String Field Theory In this section we will begin by introducing the string field products. These products are required to be multilinear, graded commutative, and to satisfy a generalized type of Jacobi identity involving the BRST operator. Closely related to the string products are the string multilinear functions. While the products are necessary to write the field equations and the gauge transformations, the multilinear functions are used to write the string field action. In this section we do not prove the requisite properties of the string products, this is done in sections 7 and 8. Here we simply show, that given those requisite properties one can build a § § string field action that is consistent. In particular we show the gauge invariance of the classical action and give the algebra of gauge transformations. For the full quantum theory we show that the action satisfies the quantum master equation and give the BRST transformations. In the last subsection we observe and comment on the fact that the algebraic structure of the classical closed string field theory corresponds to a homotopy lie algebra L∞, a cousin of the homotopy associative A∞ of Stasheff [21]. This L∞ algebra is defined on the full Fock complex of the theory, which includes physical, unphysical and trivial states. In this algebra even the ordinary Jacobi identity is violated. We also study what happens to the algebraic structure of the theory when we shift the string field by a term that does not satisfy the classical equations of motion. The resulting theory simulates closed string field theory formulated around non-conformal field theories and may be of interest in the problem of conformal background independence.

4.1. The String Products and Their Identities

Let us denote by the Hilbert space of the combined conformal field theory of matter and H ghosts, and let B1, B2, be arbitrary string fields (thought as kets). By arbitrary we mean ··· − − that while the states Bi are annihilated by b0 and L0 , they can have arbitrary ghost number G, arbitrary statistics, and need not satisfy any reality condition. The identities we will discuss hold for arbitrary string fields, but at some points we will restrict ourselves to dynamical string fields, denoted as Ψ, which are always grassmann even but may have arbitrary ghost number (in the master action), and to gauge parameters, denoted as Λ, which are Grassmann odd and of ghost number one. The general string product is denoted by B , B , , B . (4.1) 1 2 ··· n g   It has n entries, that is n states in , and is labeled by a subscript g, which is related to the H genus of the surfaces that are used to define the product. The product takes the n input states in and gives us a state in . The resulting state satisfies the subsidiary conditions H H (b b ) B , B , , B =(L L ) B , B , , B =0. (4.2) 0 − 0 1 2 ··· n g 0 − 0 1 2 ··· n g     A product defines a multilinear map from the tensor product = ⊗n to : H⊗···⊗H H H B , , B b + B′ b′ , , B = B , , B , , B b ( )b(Bi+1+···+Bn) 1 ··· i i ··· n g 1 ··· i ··· n g −     ′ (4.3) + B , , B′ , , B b′ ( )b (Bi+1+···+Bn), 1 ··· i ··· n g −  

30 for any i, such that 1 i n, and where b and b′ are c-numbers, possibly anticommuting, ≤ ≤ which is the reason we must include the sign factor taking into account that they must be moved across the B’s that appear to their right. The products are graded-commutative, that is B , , B , B , , B =( )BiBi+1 B , , B , B , , B . (4.4) 1 ··· i i+1 ··· n g − 1 ··· i+1 i ··· n g     We will define string products for all values of g 0 and for all n 0. For n =0a ≥ ≥ ‘product’ is not really a product (there are no input elements) but simply gives us a special element in . They will be denoted by [ ] , and sometimes by F , that is H · g (g) For n =0, F . (4.5) · g ≡ (g) ∈ H   For g 1 these are the states associated to surfaces of genus g with one puncture, as will be ≥ defined in 7 . For g = 0, however, we take it to be zero § 0. (4.6) · 0 ≡   For n = 1 a product takes one string field and gives us another, it is simply a linear map. When g = 0 this linear map will be simply given by the BRST operator

B 0 = QB. (4.7)   All the higher products will be defined explicitly in 7 . § The ghost number of a product of string fields is given by the following expression n G( B , B , , B )=3+ (G(B ) 2), (4.8) 1 2 ··· n g i −   Xi=1 which is g independent, and holds for all the products introduced before. Thus, for string fields of ghost number two any string product gives us an element of ghost number three. The statistics of the product is intrinsically odd, that is n

ǫ[B1,···Bn] =1+ ǫBi . (4.9) Xi=1 The fundamental identity that the products satisfy is n 0= Q B , , B + ( )(B1+···+Bi−1) B , , QB , , B 1 ··· n g − 1 ··· i ··· n g   Xi=1  

+ σ(il, jk) Bi1 , , Bil , [Bj1 , , Bjk ]g2 ··· ··· g1 (4.10) gX1,g2   {il,jk},l,k 1 ′ + ( )Φs Φ , Φ , B , B , , B , 2 − s s 1 2 ··· n g−1 Xs   e and it holds for all n 0. We will call this the main identity. The second sum in the right ≥ hand side runs over all different splittings of the set 1, , n into a first group i , , i { ··· } { 1 ··· l}

31 and a second group j , , j . Two splittings are the same if their corresponding first groups { 1 ··· k} contain the same integers, regardless of their order. The following conditions should be satisfied g 0, g 0, with g + g = g, 1 ≥ 2 ≥ 1 2 l 0,k 0, with l + k = n 0, ≥ ≥ ≥ (4.11) l 1 when g =0, ≥ 1 k 2 when g =0. ≥ 2 The factor σ(il, jk) appearing in (4.10) is defined to be the sign picked up when one rearranges the sequence Q, B , B into the order B , B , Q, B , , B . The sum in the last { 1 ··· n} { i1 ··· il j1 ··· jk } term is over states in a complete basis of the Hilbert space . This sum is restricted to states −H c satisfying the L0 L0 constraint. Moreover, since Φs = b0 Φs , the sum can be restricted to − − | i | i states Φs annihilated by b0 . This happens because the conjugate of a state Φs that is not | i − e − − | i annihilated by b0 must necessarily be annihilated by b0 , then the extra b0 in the definition of Φ kills the state and the corresponding term dissapears from the sum. This identity has its | si geometrical origin in the the conditions that string vertices must satisfy so that they generate, e upon the use of Feynman rules, the complete moduli spaces of Riemann surfaces (see 5). Its § proof will be given in 8 . § A particularly important subcase of the above equation is the case of g = 0, which is relevant for the classical closed string field theory. This implies that g1 = g2 = 0 and that the last term in (4.10) is not present:

n 0= Q B , , B + ( )(B1+···Bi−1) B , , QB , , B 1 ··· n 0 − 1 ··· i ··· n 0 Xi=1     (4.12) + σ(i , j ) B , , B , [B , , B ] , l k i1 ··· il j1 ··· jk 0 0 {Xil,jk}   l≥1,k≥2 where all the conditions on the sum are explicitly stated. While the above form for the identities might be most familiar, it is convenient, for sim- plicity, to treat the BRST operator Q as generating the product (4.7). This allows one to absorb all the terms with Q into the sums. One readily finds (just checking that the signs work out) that the general identity becomes

′ 1 Φs 0= σ(il, jk) Bi1 , , Bil , [Bj1 , , Bjk ]g2 + ( ) Φs, Φs, B1, B2, , Bn , ··· ··· g1 2 − ··· g−1 g1X+g2=g   Xs   {il,jk};l,k≥0 e l+k=n≥0 (4.13) where the conditions on the sum differ from (4.11) in that now l and k can take all possible values ( 0). The identity relevant for the classical theory becomes ≥ 0= σ(i , j ) B , , B , [B , , B ] . (4.14) l k i1 ··· il j1 ··· jk 0 0 {il,jXk};l,k≥0   l+k=n≥0

32 Note that in our present setup the terms with k = 0 vanish automatically. Let us give a few examples of these identities. Consider first g = 0, and equation (4.14). For the lowest possible value of n, namely n = 0, we find

0= [ ] Q =0, (4.15) · 0 0 → · 0     upon use of (4.7) and the fact that we defined the special vector to be zero. The next · 0 case is n = 1, and we find  

0= [B] +( )B B, [ ] = QB +0= QQB, (4.16) 0 0 − · 0 0 0       which is the statement of the nilpotency of Q. The first nontrivial identity corresponds to n = 2 where one finds

0= B , B +( )B1 B , B +( )B2(1+B1) B , B +( )B1+B2 B , B , , 1 2 0 0 − 1 2 0 0 − 2 1 0 0 − 1 2 · 0 0              (4.17) whereupon simplification gives

0= Q B , B + QB , B +( )B1 B , QB . (4.18) 1 2 1 2 − 1 2       This shows that Q acts as a derivation for the product m2. For n = 3 (deleting, for brevity, the ‘0’ subscripts from the products) one finds

0= Q B , B , B + QB , B , B + B , QB , B ( )B1 + B , B , QB ( )B1+B2 1 2 3 1 2 3 1 2 3 − 1 2 3 − + B , B , B ( )B1 + B ,B , B ( )B2(1+ B1) + B , B , B (  )B3(1+B1+B2). 1 2 3 − 2 1 3 − 3 1 2 −          (4.19) It is a good check and a straighforward exercise to verify using the above identities that

Q(Q B1, B2, B3 0) = 0. In equation (4.19), the first line corresponds to the failure of Q to be a derivation for the product m3. This failure equals the failure of the product m2, in the second line, to satisfy a Jacobi identity. ⋆ A Comment on Signs. In a graded Lie algebra one has elements a, b, and a bracket [ , ]. ··· Every element has a degree denoted by ‘deg’ and the bracket satisfies

[a, b]= ( )deg a·deg b[b, a], and deg ([a, b]) = deg a + deg b. (4.20) − − The bracket is said to be skew (graded) commutative. In closed string field theory, letting the degree be given by the statistics, the products are graded commutative (see (4.4)). The closed

⋆ The following comment was prompted by a question of J. Stasheff, and obtained in collaboration with him.

33 string bracket satisfies (g = 0)

B , B =( )deg B1·deg B2 B , B , and deg ( B , B )=1+deg B + deg B , (4.21) 1 2 − 2 1 1 2 1 2       where the second relation follows from (4.9). Thus, a priori, we do not have the properties (4.20) of the bracket of a graded Lie algebra. Nevertheless if we redefine slightly the closed string bracket and the degree letting

′ ′ deg′ B 1+ ǫ(B), and B , B ( )1+deg B1 B , B , (4.22) ≡ 1 2 ≡ − 1 2 ′     the new bracket , , with the new degree, satisfy the conditions given in (4.20). The reader may verify that with  the new product, the terms in the second line in Eqn.(4.19) become the standard three terms of the Jacobi identity of a graded Lie algebra. All this indicates that if the higher products would vanish we could obtain a standard graded Lie algebra from the closed string product. This implies that the term ‘homotopy Lie algebra’ is justified. In this work we will not use the modified products in (4.22). We will work with the original products which are directly motivated by closed string theory. Let us now consider the simplest identities for the case g 1. These correspond to g = 1, ≥ n = 0 which reads 1 ′ 0= Q + ( )Φs Φ , Φ , (4.23) · 1 2 − s s 0   Xs   e and to g = 1, n = 1, which is 1 ′ 0= Q B + QB +( )B B, + ( )Φs Φ , Φ , B . (4.24) 1 1 − · 1 0 2 − s s 0         Xs   e Let us verify the consistency of (4.23). We act with Q on the left and on the right, and we must see that acting on the right gives zero. The first term on the right is annihilated by Q, we must see that the second one is also. To this end it is sufficient to show that terms of the following form vanish: ′ ′ ( )Φs QΦ , Φ , B , , B + Φ , QΦ , B , , B =0, (4.25) − s s 1 ··· n g s s 1 ··· n g Xs   Xs   e e for any product and for arbitrary B’s. One may think that the two terms cancel each other simply by symmetry properties, but this is not quite true, in fact the second term is identical to the first; using (2.83) ′ ′ Φ , QΦ , B , , B = ( )Φs QΦ , Φ , B , , B , (4.26) − s s 1 ··· n g − s s 1 ··· n g Xs   Xs   e e and (4.4) for the second step. Thus both terms are identical ′ ′ ( )Φs QΦ , Φ , B , , B = Φ , QΦ , B , , B =0, (4.27) − s s 1 ··· n g s s 1 ··· n g Xs   Xs   e e At any rate it is easiest to establish this relation in the form (4.25) since the relevant terms

34 appear in the form

′ ( )Φs Q Φ Φ + Φ Q Φ ··· − 1| si1| si2 | si1 2| si2 ··· Xs   e e which can clearly be written as

′ (Q + Q )( )Φs Φ Φ = (Q + Q ) R′ =0 ··· 1 2 − | si1| si2 ··· ··· 1 2 | 12i ··· Xs     e f since the sum of Q’s annihilate the sewing ket. This establishes (4.25) and (4.27) and therefore the consistency of (4.23). The consistency of (4.24) is checked in a similar fashion; this time, however, one needs a particular case of the following identity

′ , Φs Φs, =0, (4.28) ··· ··· g1 g2 Xs     e where the dots denote arbitrary string fields. Another related and useful relation is

′ ( )Φr ( )Φs Φ , Φ , , Φ , Φ , =0. (4.29) − − s s ··· r r ··· g Xr,s   e e These identities will be simple to establish in 7. § A particularly useful case of the identities (4.13) satisfied by the products arises when all of the elements are the same string field Ψ (grassman even). We then have that splitting n string fields into two groups of l and k string fields respectively can be done in n!/l!k! different ways, thus Eqn.(4.13) gives

′ n! l k 1 Φs n 0= Ψ , [Ψ ]g + ( ) Φs, Φs, Ψ , (4.30) l! k! 2 g1 2 − g−1 g1X+g2=g Xs l,k≥0     l+k=n≥0 e where the sign factor σ dissapeared since it is always equal to one, and where we have defined, for brevity, the object Ψn that simply denotes a set of n entries of the field Ψ

Ψn Ψ, Ψ, Ψ . (4.31) ≡ ··· n terms | {z } Note that Ψn is not a product.

35 4.2. The Multilinear String Functions

It is now convenient to use the multilinear products discussed above, and the linear inner product introduced in 2 to define functions. These functions, given a set of string fields, give § us numbers. We define

A, B , B , , B A, B , B , B . (4.32) 1 2 ··· n g ≡ 1 2 ··· n g    Given this definition we can actually use our basis of states to reverse the relation and obtain the products from the multilinear functions. We claim that the correct relation is

′ B , , B = ( )Φs Φ Φ , B , , B . (4.33) 1 ··· n g − | si· s 1 ··· n g   Xs  e We verify this by using (4.32) in the right hand side of (4.33) to obtain:

′ (2) = ( )Φs Φ R Φ c−(2) B , , B , − | sih 12| si1 0 1 ··· n g Xs   ′ e (2) = ( )Φs Φ R b−(1) Φc c−(2) B , , B , − | sih 12| 0 | si1 0 1 ··· n g (4.34) Xs   ′ = Φ Φc B , , B , | si·h s| 1 ··· n g Xs  

− where in the first step we used (2.83) and then the c0 oscillator was moved up to the reflector c − and since it annihilates Φs it deletes the b0 oscillator. In the final expression we recognize | i − the identity operator (Eqn. (2.49)), in the subspace of states annihilated by L0 . Since the string product is in this subspace we have obtained the desired result. The functions we have introduced in (4.32) are multilinear

′ ′ b(Bi+1+···+Bn) B1, , Bib + B ib , , Bn g = B1, , Bi, , Bn g b ( ) { ··· ··· } { ··· ··· } − ′ (4.35) + B , , B′ , , B b′ ( )b (Bi+1+···+Bn), { 1 ··· i ··· n}g − for any i, such that 1 i n, and where b and b′ are c-numbers, possibly anticommuting. ≤ ≤ This equation follows from (4.3) for all the arguments except the first one. For the first one − the linearity follows from the linearity of the inner product; the sign factor arises from the c0 insertion in the inner product and the fact that the multilinear products are intrinsically odd (Eqn. (4.9)). The multilinear functions are also graded-commutative

B , , B , B , , B =( )BiBi+1 B , , B , B , , B , (4.36) { 1 ··· i i+1 ··· n}g − { 1 ··· i+1 i ··· n}g a fact that will be essentially obvious in 7 . Since a multilinear function gives us a number, §

36 which is of ghost number zero, it satisfies a selection rule

n B , B , , B =0, implies (G(B ) 2)=0. (4.37) 1 2 ··· n g 6 i −  Xi=1 The statistics of the multilinear functions is intrinsically even, that is

n

ǫ{B1,···Bn} = ǫBi , (4.38) Xi=1 The lowest order multilinear functions are B , B B , B , { 1 2}g ≡h 1 2 gi B B, ,  (4.39) { }g ≡h · gi Number  to be defined in 7, {·}g ≡ § where the last definition is not just a particular case of (4.32). For g = 0 we declare

0, (4.40) {·}0 ≡ and because of (4.6) we have B = B, = 0. The multilinear functions will enable us { }0 h · 0i to write actions next.  

4.3. The Classical String Action and its Gauge Invariance

In this section we give the classical action for closed string field theory. We then give the form of the gauge transformations, and show the invariance of the action. Our treatment just amounts to streamlining of the proofs constructed in [3]. Finally, we compute the commutator of two gauge transformations. The classical string action is simply given by

∞ 1 κn S(Ψ) = Ψn . (4.41) κ2 n! { }0 Xn=2

Here κ is the closed string field . Once we recognize that Ψ, Ψ = { }0 Ψ, [Ψ] = Ψ, QΨ , the action can be written in the more familiar form h 0i h i ∞ 1 κn−2 S(Ψ) = Ψ, QΨ + Ψn . (4.42) 2h i n! { }0 Xn=3 In fact, given that we have

Ψ = Ψ, = Ψ, 0 =0, (4.43) { }0 h · 0i h i   37 and = 0 (from (4.40)), we can actually take the action to be {·}0 ∞ 1 κn CLASSICAL ACTION: S(Ψ) = Ψn . (4.44) κ2 n! { }0 nX=0

This form of the action is motivated by the higher genus action, which must receive contribu- tions from n = 1, and may or may not include contributions from n = 0. The field equations follow from (4.41) by simple variation

∞ ∞ κn−2 κn−2 δS = n δΨ, Ψn−1 , = δΨ, Ψn−1 , (4.45) n! { }0 (n 1)! 0 Xn=2 nX=2 −   and are given by

∞ ∞ κn−1 κn−1 FIELD EQUATIONS: 0 = (Ψ) Ψn = Q Ψ + Ψn . (4.46) F ≡ n! 0 | i n! 0 Xn=1   Xn=2  

The gauge transformations of the theory are given by

∞ κn GAUGE TRANSFORMATIONS: δ Ψ = Ψn, Λ , (4.47) Λ| i n! 0 Xn=0   which in expanded form read

∞ κn δ Ψ = QΛ+ Ψn, Λ . (4.48) Λ| i n! 0 Xn=1  

Let us now show the invariance of the action under the gauge transformations. We begin with Eqn. (4.44) and obtain

∞ κn−2 δ S = Ψn−1, δΨ , Λ (n 1)!{ }0 Xn=1 − ∞ ∞ κn+m−1 = Ψn, Ψm, Λ , n! m! 0 0 Xn=0 mX=0    ∞ ∞ (4.49) κn+m−1 = Ψm, Λ , Ψn , n! m! 0 0 Xn=0 mX=0   ∞ ∞ κn+m−1 = Ψm, Λ , Ψn , n! m! 0 0 Xn=0 mX=0    

38 to find upon reordering

∞ κn−1 δ S = Ψl, Λ , Ψk . (4.50) Λ l! k! 0 0 nX=0 l≥X0,k≥0 l+k=n    

To show that this vanishes we make use of (4.30) for the case when g =0

n! 0= Ψl, [Ψk] , (4.51) l! k! 0 0 l+Xk=n≥0 l,k≥0   and form the inner product with Λ to get

n! 0= Λ, Ψl, [Ψk] , (4.52) l! k! 0 0 l+Xk=n≥0 l,k≥0   and then we simply rearrange

l k l k k l Λ, Ψ , [Ψ ]0 = Λ, Ψ , Ψ = Ψ , Ψ , Λ , 0 0 0 − 0 0 (4.53)   =  Ψk , Ψ l, Λ = Ψl, Λ , Ψk , − 0 0 − 0 0         and therefore back in (4.52) we have

n! 0= Ψl, Λ , Ψk . (4.54) l! k! 0 0 l+Xk=n≥0 l,k≥0    

This identity, holding for all n 0 implies that δ S in (4.50) is zero, as we wanted to show. ≥ Λ Gauge Algebra. Let us now turn to the algebra of gauge transformations (see also Ref.[35]). We will see that the commutator of two gauge transformations is a gauge transformation with field dependent parameter plus some extra terms which vanish upon use of the equations of motion. Consider therefore the commutator

δ , δ Ψ = δ δ δ δ Ψ , (4.55) Λ2 Λ1 | i Λ2 Λ1 − Λ1 Λ2 | i    a short computation gives us

κn+1 δ , δ Ψ = Ψl, Λ , Λ , Ψk (Λ Λ ) . (4.56) Λ2 Λ1 | i l! k!  1 2 0 0 − 1 ↔ 2  Xn=0 l,kX≥0   l+k=n    

From the basic identity for string products (4.14), consider the relation that follows when we

39 take the string of fields ΨnΛ Λ with n 0. We get 1 2 ≥ n! 0= Ψl, Λ , Λ , Ψk Ψl, Λ , Λ , Ψk l! k! 1 2 0 0 − 2 1 0 0 l,kX≥0         l+k=n≥0 (4.57) Ψl, Λ , Λ , Ψk + Ψl, Λ , Λ , Ψk , − 1 2 0 0 1 2 0 0         where the combinatorial factor is clearly the same for all four terms, and the signs arise from the oddness of the gauge parameters. This equation holds for all n 0. The first two terms ≥ appearing in (4.57) are precisely the ones that appear in the right hand side of (4.56), which therefore reduces to

κk+l+1 δ , δ Ψ = Ψl, Λ , Λ , Ψk Ψl, Λ , Λ , Ψk . (4.58) Λ2 Λ1 | i l! k!  1 2 0 0 − 1 2 0 0   l,kX≥0        

It is convenient to write this right hand side as

∞ ∞ κl κk+1 Ψl, Λ , Λ , Ψk + κ2 Ψl, Λ , Λ , (Ψ) , (4.59) l!  k! 1 2 0 0 2 1 F 0  Xl=0  Xk=0      where , the field equation, was given in (4.46). Recognizing that the first term in the above F equation corresponds to a gauge transformation (see (4.47)) we finally obtain

∞ κl+2 δ , δ Ψ = δ Ψ + Ψl, Λ , Λ , (Ψ) , (4.60) Λ2 Λ1 | i Λ(Ψ)| i l! 2 1 F 0   Xl=0   where the field dependent parameter Λ(Ψ) is given by

∞ κn+1 Λ(Ψ) = Λ , Λ , Ψn , n! 1 2 0 Xn=0   (4.61) = κ Λ , Λ + κ2 Λ , Λ , Ψ + , 1 2 0 1 2 0 ···     which is indeed field dependent beyond the first term. The second term in (4.60) is the term that vanishes when one uses the field equation (Ψ) = 0. One sees that both the field F dependent structure constants and on-shell closure of the gauge algebra are a consequence of the theory not being cubic.

40 4.4. The Quantum String Action and The Master Equation

In this section we show that the identities satisfied by the string field products and multi- linear functions imply that the full quantum action, which will be given shortly, satisfies the Batalin-Vilkovisky master equation. This implies that the string field action can be consistently quantized (in the path integral formulation). Then we will give the master transformations and the BRST transformations of the action. The BV master equation was used earlier by Hata in the analysis of the quantum HIKKO closed string field theory [39]. Before starting, however, let us derive an identity satisfied by the multilinear products which will be necessary in checking the master equation. Consider (4.30) for the case when we have a set of n 1 string fields Ψn−1, with n 1. The identity reads − ≥ ′ (n 1)! n1−1 n2 1 Φs n−1 0= − Ψ , [Ψ ]g2 g + ( ) Φs, Φs, Ψ g−1, (4.62) (n1 1)! n2! 1 2 − g1X+g2=g −   Xs   n1≥1,n2≥0 e n1+n2=n≥1 where n1 and n2 are treated somewhat asymmetrically. We now form the inner product with another string field Ψ to obtain

′ (n 1)! n1 n2 1 Φs n 0= − Ψ , [Ψ ]g2 g + ( ) Φs, Φs, Ψ g−1. (4.63) (n1 1)! n2! 1 2 − g1X+g2=g −  Xs  n1≥1,n2≥0 e n1+n2=n≥1

While n1 and n2 now appear more symmetrically in the multilinear products the prefactor with factorials is quite asymmetric. It is therefore useful to observe that

n1 n2 n2 n1 n2 n1 Ψ , [Ψ ]g = [Ψ ]g , Ψ = [Ψ ]g , [Ψ ]g , 2 g1 2 g1 2 1   n1 n2 n1 n2 = [Ψ ]g , [Ψ ]g = [Ψ ]g , Ψ , (4.64) 1 2 1 g2 n2 n1  = Ψ , [Ψ ]g , 1 g2  and this symmetry under the exchange of the pairs (n1,g1) and (n2,g2) allows us to write the first term in (4.63) as

1 (n 1)! n1 n2 n2 n1 − Ψ , [Ψ ]g2 g + Ψ , [Ψ ]g1 g , 2 (n1 1)! n2! 1 2  g1X+g2=g −   n1≥1,n2≥0 n1+n2=n≥1

1 (n 1)! (n 1)! n1 n2 = − + − Ψ , [Ψ ]g2 g , 2 (n1 1)! n2! (n2 1)! n1!  1 (4.65) g1X+g2=g − g1X+g2=g −  n1≥1,n2≥0 n1≥0,n2≥1 n1+n2=n≥1 n1+n2=n≥1

1 n! n1 n2 = Ψ , [Ψ ]g2 g , 2 n1! n2! 1 g1X+g2=g  n1+n2=n≥1 n1,n2≥0 where in the last step one breaks each sum into a piece with n , n 1, plus one term; the two 1 2 ≥ resulting sums simplify neatly and the extra two terms can be absorbed into the expression to

41 obtain the quoted result (note that n1 = n2 = 0 is not possible). Using this back in (4.63) we obtain the final form of the desired identity

′ n! n1 n2 Φs n 0= Ψ , [Ψ ]g2 g + ( ) Φs, Φs, Ψ g−1. (4.66) n1! n2! 1 − g1X+g2=g  Xs  n1+n2=n≥1 e n1,n2≥0

We now claim that the full quantum master action is simply given by the obvious gener- alization of the classical action 1 κn QUANTUM MASTER ACTION : S(Ψ) = (¯hκ2)g Ψn . (4.67) κ2 n! g Xg≥0 nX≥0 

Note that we have included, for generality, the terms with g = 0; n =0, 1, which are zero. We have also included, for g 1 the terms with n = 0, which are simply constants, and therefore ≥ do not enter into the Batalin-Vilkovisky master equation. As discussed in the introduction, all such terms may play a role in understanding background independence. The way we have defined the coupling constants is consistent since

1 κn S(Ψ) = (¯hκ2)g−1 Ψn , (4.68) h¯ n! g Xg≥0 Xn≥0  where it is clear that we have a single coupling (¯hκ2) appearing in the theory once we let Ψ k−1Ψ. Let us now verify the master equation, which reads →

∂rS ∂lS ∂r ∂lS s ∗ +¯h s ∗ =0. (4.69) ∂ψ ∂ψs ∂ψ ∂ψs

To this end it is convenient to define 1 Sn(Ψ) Ψn , S(Ψ) = h¯g κn+2g−2Sn(Ψ). (4.70) g ≡ n! g → g  g,nX≥0

Plugging this power expansion of the action into the BV equation we find

n1 n2 n ∂rS ∂lS ∂ ∂lS g1 g2 + r g−1 =0. (4.71) ∂ψs ∂ψ∗ ∂ψs ∂ψ∗ g1X+g2=g s s n1+n2=n≥0 n1,n2≥0

Note that for n = 0, which implies n1 = n2 = 0, both terms above vanish, since the corre- sponding S’s have no field/antifield dependence at all. For n = 1, the second term vanishes, and, since either n1 or n2 must be zero, the first term vanishes too. For n = 2 the selection rules (3.42) derived in 3.3 imply both terms must again vanish. Thus the relation above holds §

42 trivially for n 2. We must show it holds for n 3. It is then convenient to relabel the ≤ ≥ indices appearing above; we must show that

n1+1 n2+1 ∂ Sn+2 ∂rSg1 ∂lSg2 ∂r l g−1 s ∗ + s ∗ =0. (4.72) ∂ψ ∂ψs ∂ψ ∂ψs g1+Xg2=g≥0 n1+n2=n≥1 n1,n2≥0

In order to be able to take the derivatives from the right or from the left correctly we use the fact (from 7 ) that multilinear products are of the form B B Ω B B § { 1 ··· n}∼h | 1i···| ni with Ω grassmann even. We then have h | n n ∂rS 1 ∂lS 1 g = Ψn−1, Φ , and g = Ψn−1, Φ ( )Φs+1, (4.73) ∂ψs (n 1)! s g ∂ψ∗ (n 1)! s g −  s  − − e where the first result has no sign factor because the derivative, coming from the right hits directly the field; in the second case the derivative must be pushed across the state and we get ∂l Φs+1 ∂l the quoted sign factor. (One has ∗ Φs =( ) Φs ∗ since the derivative has the same ∂ψs | i − | i ∂ψs statistics as the corresponding antifield ψ∗, which in turn has the same statistics as the state e s e Φ , given that the string field is grassman even). | si e Let us begin the calculation with the first term of (4.72). We symmetrize explicitly in the first and second labels and using (4.73) we find

′ 1 1 n1 n2 n2 n1 Φs+1 Ψ , Φs Ψ , Φs + Ψ , Φs Ψ , Φs ( ) . 2 n ! n ! g1 g2 g2 g1 g +g =g≥0 1 2   − 1 X2 G(ΦXs)≤2     n1+n2=n≥1 e e n1,n2≥0 (4.74) The second term in the brackets can be written using (2.83) as ′ ′ n2 n1 Φs+1 n1 n2 Φs+1 Ψ , Φs Ψ , Φs ( ) = Ψ , Φs Ψ , Φs ( ) , (4.75) g2 g1 − g1 g2 − G(ΦXs)≥3   G(ΦXs)≥3   e e e e where in the last term we commuted the multilinear functions (this produces no extra sign since the string field is even and Φ and Φ are of opposite statistics). Back in (4.74) we now get a complete sum representing the first term in the BV equation (4.72)(this term is now denoted e by (I)) ′ 1 1 n1 n2 Φs (I) = Ψ , Φs g Ψ , Φs g ( ) . (4.76) −2 n1!n2! 1 2 − g1+Xg2=g≥0 Xs   n1+n2=n≥1 e n1,n2≥0 We now simplify the sum over states in the above expression ′ ′ n1 n2 Φs n1 n2 Φs Ψ , Φs Ψ , Φs ( ) = Φs, Ψ Φs, Ψ ( ) , g1 g2 − g1 g2 − Xs   Xs   e ′ e (4.77) n1 n2 Φs = Φs, Ψ Φs, Ψ ( ) . g1 g2 − Xs     e

43 In order to proceed further we use the explicit representation of the inner product

′ ′ ′ −(2) n1 (2) −(2 ) n2 (2 ) Φs = R12 Φs 1c Ψ R1′2′ Φs 1′ c Ψ ( ) , h | i 0 g1 h | i 0 g2 − Xs     ′ ′ e ′ −(2) n1 (2) −(2 ) n2 (2 ) Φs = R12 c Ψ R1′2′ c Ψ Φs 1 Φs 1′ ( ) , (4.78) h | 0 g1 h | 0 g2 | i | i − Xs     ′ ′ ′ e −(2) n1 (2) −(2 ) n2 (2 ) −(1 ) = R12 c Ψ R1′2′ c Ψ b 1′ R11′ , h | 0 g1 h | 0 g2 0 P | i     −(1′) where we moved the states to the right and recognized the sewing ket. We take b0 into ′ −(2 ) − the bra R1′2′ where it becomes b0 and then eliminates the c0 insertion. Moreover the h | n − projector 1′ is not necessary since the state Ψ 2 is annihilated by L . Thus we find P g2 0   ′ −(2) n1 (2) n2 (2 ) = R12 c Ψ R1′2′ Ψ R11′ , h | 0 g1 h | g2 | i     −(2) n1 (2) n2 (1) = R12 c Ψ Ψ , h | 0 g1 g2 (4.79) = Ψn2 , Ψn1  ,  g2 g1 = Ψn1 ,Ψn2  , g2 g1    and therefore the first term entering the BV equation (Eqn. (4.76)) is simply given by

1 1 n1 n2 (I) = Ψ , Ψ g g . (4.80) −2 n1! n2! 2 1 g1+Xg2=g≥0    n1+n2=n≥1 n1,n2≥0

The second term entering the BV equation is readily computed

n+2 ∂ ∂lS 1 ∂ r g−1 = r Ψn+1, Φ ( )Φs+1, ∂ψs ∂ψ∗ (n + 1)! ∂ψs s g−1 − s  1 ′ e (4.81) = Ψn, Φ , Φ ( )Φs+1, n! s s g−1 − G(ΦXs)≤2  e where we got no extra sign factor in the last step because the statistics of the derivative, which equals the statistics of ψ is opposite to the statistics of the state Φ . Now, as we have done s | si several times by now, the term inside the sum is written as two terms: one half of itself plus e one half of itself; one of these terms is rewritten using (2.83), we then obtain the complete sum over all ghost numbers. Thus finally, we have

1 ′ (II) = Ψn, Φ , Φ ( )Φs . (4.82) −2 n! s s g−1 − Xs  · e Now putting all together (Eqns. (4.80) and (4.82)) we have that the BV equation requires

44 ′ 1 1 n1 n2 1 n Φs (I)+(II) = Ψ , Ψ g g Ψ , Φs, Φs g−1( ) =0, −2 n1! n2! 2 1 − 2 n! − g1+Xg2=g≥0    · Xs  n1+n2=n≥1 e n1,n2≥0 (4.83) which indeed is true on account of Eqn. (4.66). This concludes our verification that the master action satisfies the quantum BV equation.

BRST transformations. In the BV formalism once we have the full master action we can immediately obtain the master transformations. These transformations become the BRST transformations upon gauge fixing. What we want to show now is that the master transfor- mation for the string field takes the form s Ψ = (Equation of Motion of Ψ) µ, that is, they | i · are proportional to the field equations arising from the master action. This fact is familiar for the classical field theory, as shown in [3], and was also noted in open string field theory [25]. We begin our derivation using Eqn.(3.28)

s Ψ = Ψ ,S µ, | i {| i }· ∂r Ψ ∂lS ∂r Ψ ∂lS = | si ∗ | ∗i s µ,  ∂ψ ∂ψs − ∂ψs ∂ψ  · ′ ∂ S ∂ S l l (4.84) = Φs ∗ Φs s µ, | i∂ψs −| i∂ψ  · G(ΦXs)≤2 e n n g n+2g−2 ′ ∂lSg ∂lSg = h¯ κ Φs ∗ Φs s µ, | i ∂ψs −| i ∂ψ  · g≥X0,n≥1 G(ΦXs)≤2 e where in the last step we made use of (4.70). Evaluating the derivatives (as in (4.73))

′ ( )Φs s Ψ = h¯gκn+2g−2 − Φ Ψn−1, Φ + Φ Ψn−1, Φ µ, | i − (n 1)!| si s g | si s g · g≥X0,n≥1 G(ΦXs)≤2   − e e ′ g n+2g−2 1 Φs n−1 = h¯ κ ( ) Φs Φs, Ψ , (4.85) − (n 1)! − | i g g≥X0,n≥1 Xs  − e 1 = h¯gκn+2g−2 Ψn−1 µ, − (n 1)! g · g≥X0,n≥1 −   where we made use of the rearrangement property (2.83), and of (4.33). The minus sign can be clearly absorbed into the definition of the anticommuting parameter µ, and the sum over n can be shifted to obtain 1 MASTER TRANSFORMATIONS: s Ψ = h¯gκn+2g−1 Ψn µ. (4.86) | i n! g · g,nX≥0  

45 This is the desired result. Indeed, for genus zero we have 1 s Ψ = κn−1 Ψn µ = (Ψ) µ, (4.87) | i n! 0 · F · nX≥0   where is the classical field equation. F In the BV formalism one obtains the BRST transformations as follows. One requires that the antifields be specified by a relation of the form ∂Υ ψ∗ = , (4.88) s ∂ψs where Υ is called the gauge . The gauge fixed path integral is given by integrating only s ∗ over fields, and substituting on the master action S(ψ , ψs ) the antifields by use of (4.88):

s 1 s s ZΥ = dψ exp S(ψ , ∂Υ/∂ψ ) . (4.89) Z −h¯  One can verify that this gauge fixed action is independent of the choice of gauge fermion Υ by use of the master equation [11]. The BRST transformations of this gauge fixed theory are simply given by the master transformations evaluated at the gauge fixing surface

∗ BRST TRANSFORMATIONS: δB Ψ (ψ = 0) (s Ψ ) , (4.90) | i ≡ | i ∗ ψ =∂Υ/∂ψ and they are only defined for the fields (that is the reason for writing (ψ∗ = 0) in the left hand side). That means that the terms in the right hand side that could be identified as transformations of antifields are simply ignored. The main property of this transformation is that it leaves the path integral in (4.89) invariant. Neither the action nor the measure are invariant but their variations cancel each other. The BRST transformations are nilpotent on the mass-shell (see, for example, [25]). + ⋆ The most well-known gauge condition is the Siegel gauge, in which one requires b0 Ψ = 0. + | i For a string field annihilated by b0 the kinetic term boils down to 1 S = Ψ c−c+(L + L ) Ψ , (4.91) kin 2 h | 0 0 0 0 | i and the propagator takes the form − + b0 b0 R12 , (4.92) L0 + L0 P| i where the operators acting on the reflector can refer to either the first or the second Hilbert space. When the propagator is given a sewing interpretation we use a proper time τ [ , 0] ∈ −∞ ⋆ One must do some work in order to obtain this gauge choice using a gauge fermion (see [25] for the case of open strings). It would be interesting to know if the Siegel gauge can always be reached, regardless of the conformal field theory and the particular details of its cohomology. The case of two dimensional strings, + where the semirelative cohomology includes states that are not annihilated by b0 might be a good case study.

46 and a twist angle θ [0, 2π]. Then b = ib− and b = b+ and we find that the propagator can ∈ θ 0 τ 0 be rewritten as 0 2π

1 (τ+iθ)L0 (τ−iθ)L0 bθbτ dτ dθe e R12 . (4.93) 2πi Z Z ·| i −∞ 0

4.5. Homotopy Lie Algebras and Axioms of Closed String Field Theory

The purpose of the present section is two-fold. We will first describe briefly a mathematical framework for homotopy Lie algebras and then discuss the algebraic structure of the classical closed string field theory. Closed string field theory formulated around a conformal background corresponds to an L algebra with products m , m , , where m denotes a product whose ∞ 1 2 ··· n input is n string fields. The stability of this algebraic structure under shifts of the string field that correspond to classical solutions was shown by Sen [27]. Here we will consider shifts of the string field that do not correspond to classical solutions. The gauge invariance of the resulting theory corresponds to a homotopy Lie algebra with products m , m , m ; where m is a 0 1 2 ··· 0 ‘product’ that, without an input string field, gives us a string field ( m0 is just a special state). We give the lowest order identities and comment on their relevance to background independence.

Homotopy associative algebras or A∞ algebras were introduced by Stasheff [21]. These † algebras, which are relevant to classical open string field theory, were used in homotopy theory. The main question was to understand when a given space X was homotopy equivalent to ‡ some loop space ΩY . It was known that an H space with a strictly associative product was homotopy equivalent to a loop space, but strict associativity, which cannot always be achieved, turned out not to be essential. Homotopy associativity, that is, associativity up to homotopy is. The result was that X is homotopy equivalent to a loop space ΩY if and only if X is and A∞ space and π0(X) is a group. An A∞ space is a space with a set of multiplications m , n 2, satisfying a set of associativity conditions. An A algebra is a similar structure n ≥ ∞ on a . Its definitions is reviewed in the work of Getzler and Jones [41]. The case when the multiplications are (graded) commutative, which leads to the homotopy Lie algebras, seems to be less familiar in the literature. We could simply take the the homotopy Lie algebra to be defined by a set of multilinear graded commutative products satisfying relation (4.14). Alternatively, we can define the homotopy Lie algebra L∞ mimicking the definition of the A∞ algebra given in Ref. [41]. This is what we do next.

Witten’s open string field theory is strictly associative [10], but if one uses open string stubs, which are † necessary to make the closed string sector explicit [40], the algebra of open string fields is only homotopy associative. A space X is an H-space if it has a multiplication m : X X X, such that for some point e one has ‡ ex = xe = x. × →

47 Given a graded vector space we consider the vector space T( ) formed by adding H H symmetrized products of H ∞ T( )= S ⊗n. (4.94) H H Xn=0 The elements of this symmetrized space will be linear combinations of elements denoted as (B , B , B ) S ⊗n. Since we take symmetrized products, the B’s above can be inter- 1 2 ··· n ∈ H changed with sign factors directly correlated to the statistics (as in the multilinear products). The space T( ) forms a cotensor coalgebra. That is, it has a comultiplication ∆ such that H ∆ : T T T. The comultiplication must be coassociative, that is, the following diagram → ⊗ must commute T ∆ T T −→ ⊗ ∆ ∆⊗1 (4.95)    1⊗∆  T y T T Ty T ⊗ −→ ⊗ ⊗ For our case comultiplication is defined as follows

∆(B , B )= σ(i , j )(B , B ) (B , , B ), (4.96) 1 ··· n l k i1 ··· il ⊗ j1 ··· jk {il,jXk};l,k≥0 l+k=n b where the sum extends over inequivalent splittings (as defined below (4.10)) and the sign factor σ is the sign picked up rearranging the sequence that appears in the right hand side into the order that appears in the left hand side. One can verify that this coderivation is coassociative. Theb extra ingredient to have a homotopy Lie algebra is a coderivation b, with b : T T, → of degree ( 1) and such that b2 = 0. A coderivation satisfies a co-Leibnitz rule, that is, the − following diagram commutes T b T −→ ∆ ∆ (4.97)    1⊗b+b⊗1  T y T T y T ⊗ −→ ⊗ We want to use the multilinear products to define the coderivation. In order to have a coderiva- tion of degree 1, we will define the degree d of B as − i d(B )=2 G(B ), (4.98) i − i then, making use of Eqn. (4.8) we find

n d( B , B , , B )=2 G( B , B , , B )= 1+ d(B ), (4.99) 1 2 ··· n − 1 2 ··· n − i     Xi=1 and now all of our products act as maps of intrinsic degree 1. The BRST operator indeed − increases ghost number by one, and therefore lowers the degree by one unit. The special

48 element must be of degree ( 1), that is, of ghost number (+3). The coderivation acts · 0 − linearly on the vectors in T, and is defined (on elements of T) as follows:

b(B , B )= σ(i , j ) B , , B , B , , B , (4.100) 1 ··· n l k i1 ··· il j1 ··· jk 0 {il,jXk};l,k≥0    l+k=n where σ is the same sign factor as that in (4.10). Note that b acting on an element in S ⊗n H gives us a sum of terms in S ⊗k with 1 k n. Rather than prove here that this is a H ≤ ≤ coderivation and that it satisfies b2 = 0 let us illustrate how the conditions b2 = 0, reproduce the lower identities we had before (taking 0 = 0, for simplicity). The simplest case is 2 · b(B1)=( B1 0)=(QB1), and therefore b (B 1)= b(QB1)=(QQB1) = 0, by nilpotency of Q. Next, we have b(B , B )= B , B +(QB , B )+( )B1 (B , QB ), (4.101) 1 2 1 2 0 1 2 − 1 2   and then if we apply b again we have

b2(B , B )= Q B , B + b(QB , B )+( )B1 b(B , QB ), (4.102) 1 2 1 2 0 1 2 − 1 2   using (4.101) to evaluate the two last terms in the right hand side we get four terms, two of which cancel, and therefore we find

b2(B , B )= Q B , B + QB , B +( )B1 B , QB =0, (4.103) 1 2 1 2 0 1 2 0 − 1 2 0       which is the familiar condition that Q is a derivation of the product m2. Shifting String Field Theory into a Non-Conformal Background. Closed string field theory is currently formulated by first choosing a conformal background, which defines a BRST operator and to some degree the string field, since the basis of vectors that we use to expand the string field is the basis of states of the conformal field theory. The algebraic structure of such a classical theory is determined by the BV equation and has been seen to be an L∞ algebra with products m , with n 1. If the string field is shifted by a classical solution, one expects the n ≥ resulting string field theory to correspond to a theory formulated around another conformal background. Indeed, upon such shift, one finds an identical algebraic structure [27] holding for modified products, which is evidence that the new theory corresponds to another conformal background. If we shift the string field Ψ by letting Ψ Ψ +Ψ′ with Ψ a string field that does → 0 0 not solve the classical equations of motion, then the new string field theory, whose fluctuation field is Ψ′, and whose vacuum corresponds to Ψ′ = 0, is a string field theory formulated around some sort of background that is not conformal. What we do next is elucidate the algebraic structure of such theory. It is simple to see that such theory still has gauge invariance. If S(Ψ) is invariant under Ψ Ψ+ δ Ψ (where δ Ψ is defined in (4.47)), then clearly S(Ψ +Ψ′) is invariant under → Λ Λ 0 Ψ +Ψ′ Ψ +Ψ′ + δ (Ψ +Ψ′), and as a consequence it is invariant under Ψ′ Ψ′ + 0 → 0 Λ 0 →

49 ′ δΛ(Ψ0 +Ψ ). This is the gauge invariance of the new action. Let us therefore begin by finding out what happens with the action. From the expression given in (4.44) we get

∞ n n 1 κ n m S(Ψ +Ψ′)= Ψ′ Ψn−m 0 κ2 n! m 0 0 Xn=0 mX=0  ∞ ∞ n 1 κ m = Ψ′ Ψn−m κ2 m!(n m)! 0 0 mX=0 nX=m −  (4.104) ∞ ∞ n+m 1 κ m = Ψ′ Ψn κ2 m!n! 0 0 mX=0 nX=0  ∞ n ∞ m 1 κ κ n = Ψ′ Ψm . κ2 n! m! 0 0 Xn=0 mX=0  This result suggests that we define a new set of multilinear functions and products, denoted by primes ∞ m ′ κ B , , B = B , , B , Ψm , 1 ··· n m! 1 ··· n 0 mX=0   (4.105) ∞ m ′ κ B , , B = B , , B , Ψm , 1 ··· n m! 1 ··· n 0   mX=0   which are therefore related to each other as ′ ′ B , B , , B = B , B , B . (4.106) 0 1 ··· n h 0 1 ··· n i    Using the definition of the primed multilinear functions, the shifted action in Eqn. (4.104) simply reads ∞ 1 κn S(Ψ +Ψ′) S′(Ψ′)= Ψn ′. (4.107) 0 ≡ κ2 n! { } Xn=0 The new action therefore takes the same form as the original action (in (4.44)) with the multilinear functions changed into new ones. The first two terms in the expansion are

∞ m 0 κ Ψ′ = Ψm κ2S(Ψ ), { } m! { 0 } ≡ 0 mX=2 (4.108) ∞ m ∞ m 1 κ κ Ψ′ = Ψ′, Ψm = Ψ′, Ψm = Ψ′, κ (Ψ ) , { } m! { 0 } h m! 0 i h F 0 i mX=1 mX=1   where was defined in (4.46). Thus the shifted action, reads F S(Ψ +Ψ′) S′(Ψ′)= S(Ψ )+ Ψ′, (Ψ ) + , (4.109) 0 ≡ 0 h F 0 i ··· which, of course, is the expected form of the expansion, since hadΨ0 satisfied the field equation, the term linear in the fluctuation field Ψ′ would have vanished. Let us now consider the gauge

50 transformations. As we discussed above the shifted action is invariant under the transformation

∞ κn δ′ Ψ′ = δ (Ψ +Ψ′)= (Ψ +Ψ′)n, Λ , (4.110) Λ Λ 0 n! 0 nX=0   expanding out and rearranging in the same way as we did for the action above, one finds

∞ n κ n ′ δ′ Ψ′ = Ψ′ , Λ , (4.111) Λ n! Xn=0   which is just the same form of the standard gauge trasformations but now with the primed products. The first term corresponds to the new BRST-like operator

∞ n ∞ n ′ κ κ Λ Q′Λ= Λ, Ψn = QΛ+ Λ, Ψn . (4.112) ≡ n! 0 n! 0   nX=0   Xn=1   As we will see this new BRST-like operator does not quite have the standard properties. The reader may wonder how it is possible that the structures defining the action and gauge transformations appear not to have changed despite the shift that did not satisfy the field equations. Actually the structures have changed! When we began with a conformal theory we had products m , m , but for convenience our whole analysis was done including the 1 2 ··· product m0, which was set to zero, and the associated multilinear function, that was also set to zero. Our proof of gauge invariance, beginning in (4.47) never assumed that the product m0 (and the associated multilinear function) was zero, we simply used the basic identity (4.14) for all values of n. This time m0 is not zero anymore. Nevertheless the gauge invariance of the shifted action under the shifted gauge transformations implies that the primed products must satisfy (4.14)! Thus, the structure of string field theory formulated around a background that is not conformal is that of an L∞ algebra with products beginning with m0. It is fun to see what happens with the identities we are so familiar with. This time we will denote the special element defined by m0 by the name F ,

′ F. (4.113) · ≡   The state F must be Grassmann odd and of ghost number (+3). Note that, from Eqn. (4.108) we have that 1 ′ Ψ′ Ψ′, = Ψ′, κ (Ψ ) , (4.114) { }≡h · i h F 0 i   and therefore, for our previous example F = κ (Ψ ). The lowest identity in (4.14) was F 0 analyzed earlier in (4.15) where we had

′ ′ 0= [ ]′ Q′ = Q′F =0. (4.115) · → ·     That is, the special element is annihilated by the new ‘BRST’-like operator. This is a nontrivial

51 identity. The next case is that in (4.16) which reads

′ ′ ′ ′ 0= [B]′ +( )B B, [ ]′ = Q′B +( )B B, F =0, (4.116) − · −         which implies that 2 ′ Q′ B + F, B =0, (4.117)   and therefore the new BRST-like operator fails to be nilpotent by a term involving the product m2 and F . The next identity is (4.17) which this time gives

′ ′ ′ ′ 0= Q′ B , B + Q′B , B +( )B1 B , Q′B + F, B , B , (4.118) 1 2 1 2 − 1 2 1 2         that is, Q′ also fails to be a derivation by a term involving F . It seems that the above discussion could be useful for our understanding of background independence. The analogy with Einstein’s gravity is interesting. In this theory, the metric on a is the dynamical variable, and the Ricci-flat metrics are classical solutions. The analogous objects in string theory are not clearly known, but it would seem reasonable to believe that the dynamical variable (string field) in some sense specifies a two-dimensional field theory, and the classical solutions are the conformal field theories. Thus our present knowledge is as if we knew how to write Einstein’s equations around Ricci-flat metrics, but not around arbitrary metrics. If we were in this situation, by shifting the metric a bit we would be simulating Einstein’s equations around a non-Ricci-flat metric. This is what we did above for string theory. In this sense our problem in string theory is not that we use a two- dimensional field theory to write the string action, it is that we use a conformal field theory. If we knew how to write the string action for a more general class of theories it would be an advance. What we did was exploring indirectly how such theory would look. It would be interesting to study more directly, in two-dimesional field theory, what is the operator Q′, and what is the state F . As a last point I wish to emphasize that gauge invariance of the Einstein action expanded around an arbitrary background works in a way exactly analogous as that discussed above for closed string theory. Consider the Einstein action, and an arbitrary background metric gαβ that does not satisfy the field equations. Einstein’s action for the metric gαβ = gαβ + hαβ would read

1 4 1 4 µν 1 µν 2 S(g)= g Rd x + d x g R g R hµν + (h ), (4.119) −κ2 Z κ2 Z − 2 O p p   Now, the gauge transformations on the original metric, imply the following transformations for the fluctuating field hµν

δǫhµν = Dµǫν + Dνǫµ, (4.120)

The above expansion of the action in powers of the fluctuation is gauge invariant, this requires

52 that the variation of the term linear in h be zero, which happens because

µν 1 D R gµν R =0, (4.121) ν − 2   which is simply the Bianchi identity. The last three equations are entirely analogous to our previous equations S(Ψ +Ψ′)= S(Ψ )+ Ψ′, (Ψ ) + , 0 0 h F 0 i ··· δ′ Ψ′ = Q′Λ+ , Λ ··· Q′ =0, (4.122) F which were used to study the gauge invariance of the string field theory formulated around a non-conformal background.

5. String Vertices and The Geometrical Identity

String vertices are actually defined, in most generality, by Riemann surfaces with punc- tures and coordinates around the punctures. In covariant string field theory, the vertices are universal in the sense that they may be used for any background around which the theory is formulated. Vertices are the Riemann surface ingredient of the string field theory. In fact, finding a consistent set of vertices was the main difficulty in constructing closed string field theory. In 5.1 we will define precisely what vertices are, and give the basic properties they § must have in order that the field theory defined by these vertices yield manifestly factorizable and unitary amplitudes, with correct symmetry properties and hermiticity. The conditions that eventually yield BRST invariance will be given in 5.2. § 5.1. Definition of the String Vertices

A string vertex, at a given genus g and for a process involving n strings is a set of surfaces. ⋆ This set is the subset of the full (compactified) moduli space of Riemann surfaces , Mg,n and as we will see in 6, it corresponds to the surfaces that are not produced by the sewing § procedure. The corresponding string field vertices, which define the interactions of string fields will be discussed in 7. Let us now define the string vertices. § Definition. A string vertex is a set of Riemann surfaces of genus g and n punctures repre- Vg,n senting a subset of . This set does not include surfaces arbitrarily close to degeneration: Mg,n there exist open neighborhoods of the compactification divisor of that does not Mg,n Vg,n intersect. Each surface in the set must be equipped with a specific choice of analytic local

⋆ The compact boundaryless space g,n is obtained from g,n by adding the surfaces with nodes, or degenerate surfaces. Such surfaces,M also called the compacMtification divisor, are obtained by sewing ordinary surfaces in the limit of vanishing sewing parameter.

53 coordinates around each of its punctures. The coordinates around each puncture are only de- fined up to a constant phase. The coordinates must be defined continuously (up to the phase ambiguity) over the set . Vg,n Let us elaborate on this definition. In particular we will impose further conditions on the vertices. The fact that local coordinates are defined up to phases means that two sets of ′ local coordinates zi and zi on the same punctured surface are said to be equivalent if for any ′ point P around the i-th puncture, such that zi(P ) is defined, zi(P ) is also defined, and one has z (P ) = exp(iθ )z′(P ), for all i. The θ ’s are puncture dependent constant phases. The set i i i i Vg,n may not be connected nor simply connected so continuity of the definition of local coordinates is understood to hold for each connected component of the set. Continuity also holds up to the constant phase ambiguity. It is necessary for continuity of the off-shell amplitudes as we move in moduli space. This, in turn, is required for BRST invariance. It is probably impossible to get rid of the phase ambiguity in the definition of the local coordinates in the string vertices. It is well known that one cannot define local coordinates without phase ambiguities all over moduli space because of a topological obstruction [37]. Whether one can define them continuously over the subsets is not clear, but seems rather Vg,n hard since the ’s are expected to have complicated topology. At any rate neither quadratic Vg,n differentials nor minimal area metrics show us how to avoid the phase ambiguities. Since the scale of the local coordinates must be fixed, we will impose the following require- ment: (a) For every surface the local coordinates z around each puncture must define a R ∈ Vg,n i disk D = z 1. The disks must not overlap, except possibly at their boundaries. i | i| ≤ The case when the disks Di cover precisely the complete Riemann surface, are called overlap type vertices. In general this will not be the case, and removal of the disks leaves a surface with a number of holes. While the classical closed string field theory can be constructed with overlap type vertices, this is not the case for the quantum theory, as we shall see in the next section. The complete information about the local coordinates is actually encoded in the disks † Di. A disk Di whose boundary is a simple closed Jordan curve γi (a curve without self- intersections) allows us to define local coordinates in a very simple way: we declare the curve γ to be the locus z = 1, and the puncture to correspond to z = 0. The coordinates in the i | i| i disk Di are defined by the conformal mapping to the unit disk. The three parameter ambiguity in the conformal map relating two disks is reduced to one via the constraint on the position of the puncture. The left over parameter is the ambiguity due to rigid rotations of the disk. Therefore

Definition. A punctured Riemann surface with local coordinates around the punctures pi, isa Riemann surface with simple closed Jordan curves γi, called coordinate curves, homotopic to the punctures, defining punctured disks D , such that int(D ) int(D )= 0 for i = j, where i i ∩ j 6 int(Di) denotes the interior of the disk Di.

I thank A. Connes for this observation. †

54 Our inability to define phases corresponds to the fact that we cannot single out a point in every coordinate curve continuously over moduli space. Two local coordinates agree (up to phases) if their coordinate curves are identical. While the coordinate curves are, a priori, curves without parametrization, the map to the unit disk gives them a canonical parametrization. The condition that the surfaces in be non-degenerate surfaces is essential since these Vg,n surfaces will give rise to the interaction vertices of the string field theory and we insist that all degenerate surfaces arise from propagators, in order to have manifest factorization and unitarity. It is natural to require for a covariant theory, that these vertices be symmetric under the exchange of punctures, thus, even though our punctures are always distinguishable, or labeled punctures, we demand that (b) The assignment of local coordinates to the punctured surfaces is independent of the labeling of the punctures, moreover, if a surface with labeled punctures is in then copies of with R V R all other inequivalent labelings of the punctures are also included in . V This condition has been discussed in detail in [7], where we also required that:

(c) If a surface then ∗ , where ∗ denotes the mirror image of ; moreover, the R ∈ V R ∈ V R R local coordinates in and ∗ are related by the antiholomorphic map that relates the two R R surfaces. This condition is required for manifest hermiticity of the string field interactions defined by the string vertices [7]. We now turn to the conditions that will guarantee BRST invariance of the theory.

5.2. Consistency Conditions on String Vertices

The choice of vertices is not arbitrary. There is a very stringent consistency condition represented in the following geometrical equation [7]

(5.1)

In this equation the left hand side denotes the set of surfaces in that lie on the boundary Vg,n of the region of moduli space this vertex fills. These are surfaces with labeled punctures. In the right hand side the first object is a sum of groups of terms, each such group is labeled by a choice of numbers (g1, n1) and (g2, n2), where g1 + g2 = g, and n1 + n2 = n + 2. These numbers indicate the genus and number of punctures of the surfaces we will deal with. Having done this we split the n labeled punctures into two groups, (n 1) for the surfaces having n 1 − 1 punctures, and (n 1) for the surface having n punctures. The punctures left open are to 2 − 2

55 be sewn to each other. Since this splitting can be done in in n different ways this is the n1−1 number of terms we obtain for this group. For each term we now consider the sets and Vg1,n1 , with the labeled punctures assigned and one puncture left over (the way to assign the Vg2,n2 labeled punctures is immaterial since the vertices are symmetric, as demanded in property (b) 5.1). We then sew each surface in the set to every surface in the set . This sewing § Vg1,n1 Vg2,n2 is done by fixing arbitrarily the phase ambiguity in the two punctures to be sewn (marking a point in the coordinate curves) and then sewing them according to the relation z1z2 = t, where the sewing parameter t takes all the values in the unit circle t = 1, and where z and z are | | 1 2 the local coordinates around the punctures. It is clear that n , n 1, since there must be at 1 2 ≥ least one puncture in each surface in order to sew them together. When g = 0, both n1 and n2 must be greater or equal to three. The factor of 1/2 preceding this first object in the right hand side of the equation is necessary because every surface that this first object produces is actually produced twice. This follows from the above description, the sum double counts even when g1 = g2, since then the splitting of the labeled punctures still works as if the surfaces were of different genus. The second term denotes the set of surfaces obtained by taking each surface in the set , choosing arbitrarily two punctures, fixing marked points on the coordinate curves, Vg−1,n+2 and sewing the punctures via z z = t with t taking all values in the unit circle t = 1. 1 2 | | This term only exists for g 1. A factor of 1/2 is also necessary in this case. This happens ≥ because once more surfaces are produced twice. Say we label punctures ‘one’ and ‘two’ to be the punctures to be sewn. Any surface in the set can be paired to another surface in the set V which differs only by the interchange of labels on these two punctures (this is condition (b) in 5.1). Upon sewing, the original surface and its corresponding pair will produce the same sets § of surfaces. Eqn.(5.1) applies for n 3 when g = 0. This is equivalent to setting 0, for ≥ V0,n ≡ n = 0, 1, 2. It also applies for all when g 1 with the exception of (the one-loop Vg,n ≥ V1,0 cosmological constant). Note that for this case (g =1, n = 0), there are no candidate terms in the right hand side. Thus the vacuum vertex is not constrained by (5.1). All the higher V1,0 vacuum vertices are. The equality of the sets requires that the surfaces agree both in their moduli parameters and in their local coordinates around the punctures. Both the right hand side and the left hand side are sets of equal dimensionality; the boundary operation subtracts one real dimension, and sewing adds one real dimension. The set appearing in the left hand side (without the V boundary operator) is of one complex dimension higher than that of the sets appearing on the right (without sewing). There are also signs in Eqn.(5.1) (these signs, and the symmetry factor, which were ommitted from the graphical representation in Refs.[7,5], will be included here). The boundary of in the left-hand side is some subspace of with an orientation on it. Vg,n Mg,n The orientation is induced by that of . The same subspace with an opposite orientation Mg,n would be accompanied by an opposite sign. For the terms in the right-hand side one should note that these correspond to Feynman diagrams built with one propagator in the limit when the propagator collapses. This is so because the usual propagator corresponds to sewing with parameter t in the region t 1. One may therefore fix the orientation of the terms in | | ≤

56 the right-hand side of Eqn. (5.1) by thinking of them as boundaries of the regions of Mg,n obtained with t < 1. Typically there are cancellations between terms in the right-hand side | | of Eqn.(5.1). This equation is an equation in the sense of homology, where we can add or subtract with a given orientation. The above equation follows if the string vertices , generate a single cover of each moduli V space when used together with propagators (sewing with t 1) to build the relevant Mg,n | | ≤ Feynman graphs. The basic statement in Eqn.(5.1) is that the vertex and the Feynman Vg,n graphs with one propagator cover regions lying on opposite sides of their common boundary. This explains why there is a minus sign in the right hand side of the equation. Denoting by RI the region of moduli space covered by all the Feynman graphs constructed with I internal lines, the above equation can be written as

∂ = ∂ R , (5.2) Vg,n − p 1 where ∂ denotes a boundary obtained by propagator collapse ( t = 1). If we have a single p | | cover of moduli space, the full moduli space should be obtained by the union of all the regions covered by the various Feynman graphs

= R R , (5.3) Mg,n Vg,n ∪ 1 ∪···∪ 3g−3+n where (3g 3+ n), the complex dimension of , is also the maximum possible number of − Mg,n propagators. Equation (5.2) holds for the lowest dimensional moduli spaces, those corresponding to four punctured spheres and one-punctured tori. This happens because in these two cases Eqn.(5.3), which is the assumption of single cover, reduces to = R , since there are only graphs Mg,n Vg,n∪ 1 with one propagator. The absence of boundary of implies that ∂ = ∂R . Moreover, M V − 1 since the Feynman graphs use only three point vertices, which have no moduli, the boundary ∂R1 coincides with the propagator boundary ∂pR1. Equation (5.2) is established in general by induction [7] and we elaborate on the proof next. First note that any Feynman graph with I propagators covers some region of moduli space. This region can be parameterized by the propagator parameters and the parameters that specify the vertices involved in this particular Feynman graph. For each choice of these V parameters we get a specific surface, and different choices of parameters must correspond to different surfaces, for otherwise we would have multiple cover of moduli space. This implies that the boundary of the region of moduli space that we get from this Feynman graph corresponds to the boundary of the parameter space of the graph. Thus we can have either propagator boundaries, denoted by ∂p, when a propagator collapses, or vertex boundaries, ∂v, when some vertex used in the Feynman graph has its parameters lying on the boundary of the parameter space of that vertex. Thus, for the region of moduli space produced by any Feynman graph fI with I propagators we can write

∂fI = ∂pfI + ∂vfI , (5.4) and therefore, for RI = fI , which is the total region of moduli space arising from Feynman P 57 graphs with I propagators, a similar equation holds

∂RI = ∂pRI + ∂vRI , (5.5) which says that the boundary of RI is also of two types, boundaries that can be seen to arise from some graph on a propagator boundary, and boundaries that arise from some graph on a vertex boundary. Since has no boundary, we then have Mg,n ∂ 0= ∂ + ∂ R Mg,n ≡ Vg,n p 1 +(∂vR1 + ∂pR2)+ . . (5.6)

+(∂vR3g−4+n + ∂pR3g−3+n)

+ ∂vR3g−3+n

We now claim that the induction hypothesis, namely that Eqn.(5.2) holds for any moduli space of dimension lower than that of , implies that Mg,n ∂ R = ∂ R , I 2. (5.7) p I − v I−1 ≥ We will show this next. In the left hand side, whenever a propagator collapses, we can look at the vertices involved, these may be just one, if the propagator joins two legs of a single vertex, or two, if it joins two vertices. Call this the collapsed subdiagram. It is a subdiagram since the graph must have at least another propagator (I 2). The collapsed subdiagram corresponds ′ ≥ to some ′ ′ , where n is the number of legs attached to the vertex (or vertices) without Mg ,n counting the legs involved with the collapsed propagator, and g′ is the genus of the vertex (or sum of genera of the vertices). The dimensionality of ′ ′ must be lower than that of , Mg ,n Mg,n since the graph has additional parameters that those involved in the subdiagram (at least the parameters of the extra propagator).

Consider then any given surface in ∂pRI . The collapsed subdiagram is of the form ∂pf1, that is, a propagator boundary of some Feynman graph with one propagator. Consider now keeping fixed all the parameters of the graph that do not correspond to the subdiagram. This includes the parameters of the propagators joining into the subdiagram, or the parameters of propagators that may be joining two legs of the subdiagram; such propagators are not, in general, at their boundary values. Since using Feynman rules one must include all possible diagrams, it is clear that keeping these parameters fixed, we must find in ∂pRI the diagram in question, with the collapsed subdiagram not only taking the value ∂pf1 but rather running over the whole set R associated with the moduli space ′ ′ . Therefore we can use the induction 1 Mg ,n hypothesis, Eqn.(5.2) to replace the collapsed subdiagram ∂ R by a vertex ∂ ′ ′ at a p 1 − vVg ,n vertex boundary. It should be noted that Eqn.(5.2), or its graphical representation in Eq.(5.1), can be used even if the n external states are actually not external (as is our case now), but are connected via sewing to other vertices. This is so because we can think of Eqn.(5.1) as

58 an equation relating surfaces with boundaries, given the control we have over the coordinate systems at the punctures. In this process we have obtained a diagram with one less propagator and with a vertex at a vertex boundary. This is just the statement of Eqn.(5.7). One now makes use of Eqn.(5.7) to cancel all pairs of terms in parenthesis in Eqn.(5.6). Moreover the last term vanishes since the corresponding Feynman graph is built with three point vertices only and those have no vertex boundary. It therefore follows that the first two terms in the right hand side of Eqn. (5.6) must add up to zero. This shows Eqn.(5.2) holds to next order in the dimensionality of moduli space, and therefore in general. Let us note the significance of the geometrical equation (5.1). This is a consistency condi- tion for sets of Riemann surfaces with local coordinates, and has nothing to do with a specific choice of conformal field theory, or string theory. As we will see later the existence of ’s Vg,n satisfying this condition is the main geometrical input to the full off-shell field theory of closed strings. This equation encodes the Riemann surface aspect of string theory, to all orders in perturbation theory. Given that the existence of a set of string vertices satisfying the equation has such important consequences, it is essential to make sure such consistent string vertices exist. This will be the subject of section 6, where we will use minimal area metrics to find § what appears to be the simplest possible solution of the consistency conditions. As it turns out a vertex is necessary for every moduli space. Our strategy will be to show that we have a single cover of moduli space when we choose suitable ’s. Therefore, as we discussed above, Vg,n the string vertices will satisfy the consistency conditions. Equation (5.1) may describe some interesting algebraic structure on the space of Riemann surfaces, with the subsets representing generators and the product operation being sewing. Vg,n Our use of (5.1), however, will be based on the fact that it will help us define, in 7, a useful § algebraic structure on string fields, that is, on the state space of conformal field theories.

6. Minimal Area String Diagrams

In the present section we will discuss the basic ideas concerning minimal area metrics. The main purpose is to determine a canonical set of string vertices that satisfy the consistency Vg,n condition (5.1). Our analysis will be different from that of [8] in that we will not emphasize the relation to quadratic differentials. Using the results of Ranganathan [15], Wolf and the author [16] and some further results given here we will be able to give a characterization of the string vertices which is intuitively compelling, and far more concrete than any previous Vg,n description. Moreover, we will show how to obtain the unique Feynman graph corresponding to any surface by inspection of its minimal area metric.

59 6.1. Properties of the Metric of Minimal Area

A (conformal) metric ρ on a Riemann surface defines a length element dl = ρ dz and an R | | area element ρ2dx dy. Here ρ must be positive semi-definite, and since dl2 = ρ2(dx2 + dy2), ∧ ρ2 is recognized to be simply the Weyl factor of the metric. Whenever we talk about metrics on Riemann surfaces we refer to such conformal metrics. For any surface we are interested R in the metric of least possible area under the condition that all nontrivial closed curves on R be longer than or equal to 2π [8]. This minimal area problem, in contrast with the related problems studied earlier in the mathematical literature (see [29]) does not require specifying some homotopy classes of curves on the Riemann surface on which we impose length conditions. The length conditions are imposed on all curves. This is why this is a problem defined on moduli space, and why the resulting metrics have nice factorization properties (= nice behavior on the compactifying divisor). A metric is called admissible if it satisfies the length condition

ρ dz 2π, (6.1) Z | | ≥ γ for all nontrivial closed curves γ . A useful property of the space of admissible metrics is ∈ R that it is a convex space; if ρ0 and ρ1 are admissible metrics then

ρ = (1 t)ρ + tρ , (6.2) t − 0 1 is an admissible metric for all t [0, 1], as it follows directly from the above two equations. ∈ The area functional has a very nice property: it is strictly convex. This means that for the A above one-parameter family of metrics, with ρ = ρ , one has 0 6 1 (ρ ) < (1 t) (ρ )+ t (ρ ), (6.3) A t − A 0 A 1 for t (0, 1). The inequality is strict since we do not include the endpoints t = 0, 1 corre- ∈ sponding to ρ0 and ρ1. The above relation is quite familiar, and is derived using the Schwartz inequality. It simply means that the area, as a function of t, is convex (see Fig. 1). An imme- diate consequence of this fact is the uniqueness of the minimal area metric (if it exists). This is argued as follows. Assume there are two different metrics ρ and ρ on the surface , both 0 1 R admissible and both having the (same) least possible area A. It follows from convexity of the space of admissible metrics that ρ (for any t (0, 1) ) as defined above is admissible, and due t ∈ to the convexity of the area functional (ρ ) < (1 t)A + tA = A, in contradiction with the A t − assumption that A was the least possible value for the area. The uniqueness of the minimal area metric is fundamental for our purposes. Another simple consequence of convexity of the area is that given two admissible metrics ρ and ρ such that (ρ ) > (ρ ), we then have 0 1 A 0 A 1 (ρ ) < A(ρ ) for t (0, 1) (see Fig. 1). A t 0 ∈ A metric solving the minimal area problem is expected to give rise to closed geodesics of length 2π that foliate the surface completely. The reason for this was given in [5] and elaborated

60 in [16]: through every point where the metric is continuous there must exist nontrivial closed curves of length arbitrarily close to 2π, otherwise the metric could be lowered at this point without destroying admissability. While further discussion is necessary to prove the existence of a geodesic of length precisely 2π, it seems clear intuitively that this curve must exist. One can show that such curves must foliate the surface since they cannot intersect whenever they are of the same homotopy type (see below). We shall call the length 2π geodesics by the name saturating geodesics. They can be thought to describe the closed strings on the Riemann surface. We can actually go a long way towards understanding the properties of the minimal area metric by the use of the following Lemma, suggested to me by W. Thurston [42]. Lemma 6.1. Two saturating geodesics that intersect at a point where the metric is smooth cannot intersect elsewhere. Proof. Consider two saturating geodesics and . By definition they are nontrivial simple C1 C2 closed curves of length 2π. They may or may not be homotopic to each other. Pick the intersection point p1 where the metric is smooth, and assume there are more intersection points. Find another intersection point p in such that there is a segment p p in 2 C1 1 2 C1 containing no other intersection points. Denote this segment by a′ and the remaining of by C1 a. The points p ,p also determine two segments on , they will be denoted as b′ and b (see 1 2 C2 Fig.2). The symbols a, a′, b, b′ will denote both the segments and their lengths; a b will mean ∼ that the two segments are homotopic, a >,<, = b will mean relations between their lengths. { } We have a + a′ =2π, b + b′ =2π, (6.4)

We claim now that neither a nor a′ can be homotopic to b or b′. Suppose a b′, then we have ∼ that two new nontrivial closed curves: a′ + b′ a′ + a and b + a b + b′. If b′ < a then the ∼ ∼ curve a′ + b′ would be smaller than 2π. If, on the contrary a < b′ then the curve b + a would be smaller than 2π. Therefore a = b′, and the curves a + b′ and b + a would be saturating geodesics. Their lengths, however, can be shortened at the point p1, where these geodesics have corners. This shows that a b′. The other cases are entirely analogous. 6∼ Given two open curves with identical endpoints which are not homotopic to each other, they define a nontrivial closed curve. Thus we must have four nontrivial closed curves giving the following conditions: a b a + b 2π, 6∼ → ≥ a b′ a + b′ 2π, 6∼ → ≥ (6.5) a′ b a′ + b 2π, 6∼ → ≥ a′ b′ a′ + b′ 2π. 6∼ → ≥ Adding pairs from the above equations and using (6.4) we then find that a, a′, b, b′ π, which { } ≥ implies that actually a, a′, b, b′ = π, and that the four curves above are saturating geodesics. { } Since they all have corners their lengths can all be reduced, and this is a contradiction to admissibility. This concludes our proof of the Lemma.

61 Let us now consider Riemann spheres with punctures. Let us understand why the minimal area metrics here always arise from Jenkins-Strebel quadratic differentials. Since any saturating geodesic on a genus zero surface with punctures must cut the surface into two separate pieces, if two saturating geodesics cross they must do it at least in two points. The above lemma indicates that this cannot be a generic situation, thus we are led to expect that the surface will be foliated by bands of geodesics that do not intersect. This is precisely what happens with Jenkins-Strebel quadratic differentials, the horizontal trajectories of the quadratic differential are the saturating geodesics. The surface is then foliated by bands of geodesics, the so-called ring-domains of the quadratic differential. The horizontal trajectories can intersect in the critical graph of the quadratic differential, but this graph is of zero measure on the surface. It also becomes clear from the lemma that in higher genus Riemann surfaces one can very easily have saturating bands of geodesics that cross. This already happens at genus one, where curves along the “a-cycle” and “b-cycle” are both nontrivial and intersect only once. Thus higher genus minimal area metrics do not always arise from quadratic differentials. One concrete example showing such nontrivial crossing of foliations was given in [16].

6.2. Results and Open Questions in Minimal Area Metrics

Let us try to make clear what we still do not know about minimal area metrics. Still missing is a proof of existence of such metrics. The evidence in favor of existence is very strong; the minimal area metrics are known for all Riemann spheres and for large subsets of the higher genus moduli spaces (the restricted Feynman graphs of Ref. [8]). We now have even examples of minimal area metrics that do not arise from quadratic differentials [16]. Still an abstract proof of existence is not available yet. Another important property is that of completeness of the minimal area metric. This has to do with the presence of punctures; since curves surrounding punctures must also be longer than or equal to 2π, we expect that neighborhoods of the punctures must look like tubes going off to infinity. Thus we expect that any curve reaching the puncture would be of infinite length. This is the statement of completeness of the metric. It is simple to convince oneself that an incomplete metric is innefficient as far as minimizing area, but a proof of completeness is not available yet. Finally, we need smoothness of the minimal area metric in some neighborhood of each puncture. Again, it is simple to convince oneself that discontinuous metrics are inefficient, but a proof is required. In summary, we have the following assumption Assumption. We assume the existence of a complete minimal area metric smooth in some neighborhood of each puncture. At the present time we can prove, with the above assumption, that the minimal area metrics have the requisite conditions to define the string diagrams of covariant closed string field theory. The requisite conditions were essentially those of flatness and amputation [5]. Flatness requires that there should exist a neighborhood of the punctures that it isometric to a flat semiinfinite cylinder of circumference 2π. This is established by showing first that completeness and smoothness implies that there must exist a neighborhood of each puncture which is foliated by a single band of geodesics, all homotopic to the puncture (Lemma 2.1 [16]).

62 Then one uses the result of Ranganathan [15], that in a region foliated by a single homotopy type of geodesics the metric is flat. A similar result was established in [16] where it was also shown that the neighborhood is isometric to a flat semiinfinite cylinder of circumference 2π. The basic tool for the proof of flatness in [16] was a lemma showing how to construct local deformations of admissible metrics. One begins with a metric ρ0 smooth over a domain foliated by (posibly several bands of) saturating geodesics and perturbs it into ρ0 + ǫh where h is a smooth metric with compact support in the domain and such that no saturating geodesic becomes smaller than 2π in the new metric. The new metric is not yet admissible, since the old saturating geodesics, which are still fine, may not be the shortest curves in the new metric. One can show that the addition of an ǫ2 term to the perturbed metric restores admissibility. 2 This means there is a constant K such that the metric ρǫ = ρ0 + ǫh + ǫ Kχ, where χ = ρ0 over a compact domain, is admissible. The useful consequence is that the derivative with respect to ǫ of the area of the metric ρǫ is simply controlled by the sign of ρ0h. In order to show a metric is not of minimal area one must find a suitable h which makesR the above integral negative, which implies that for sufficiently small ǫ the new admissible metric ρǫ has less area. The flatness property implies that the minimal area metric defines a canonical domain around each puncture. The metric must cease to look like a flat semiinfinite cylinder at some point. We define the curve to be the boundary of the maximal region foliated only by C0 geodesics homotopic to the puncture. We let be the saturating geodesic in the cylinder a Cl distance l away from . The maximal region is a canonical domain which will be used to C0 define the local coordinates around the punctures (see Fig.3 (a)). The amputation property (Fig.3 (b)) states that given a minimal area metric on a surface with punctures, amputation of a semiinfinite cylinder along the curve for l > 0 yields a R Cl surface with boundary whose minimal area metric (under the assumption that it be smooth in a neighborhood of the boundary) is simply the one induced by the original metric in the surface (see [16], Thm. 5.1). This property is essential to show that the plumbing of surfaces R with minimal area metrics gives a surface with a minimal area metric [5,16]. The basic idea is simple, given two surfaces to be sewn together (or a single surface with two legs to be sewn), one first amputates the semiinfinite cylinders corresponding to the relevant legs. Given the amputation property the resulting surfaces with boundaries have minimal area metrics. If we glue together the open boundaries and by doing this we do not create closed curves that are smaller than 2π (this is the reason for stubs, as we will see) then the resulting surface inherits a minimal area metric. This is so because any candidate metric with less area would have to have less area in at least one of the surfaces that were glued, but this is impossible given that the amputated surfaces have minimal area metrics. The reader interested in a rigorous argument should consult Ref. [16].

63 6.3. Heights of Foliations and Cutting

We need a couple of additional results in order to give a simple characterization of the string vertices . It is convenient to define the height of a foliation. Consider a band of Vg,n saturating geodesics that make up a foliation. Such a band corresponds to an annulus, and this annulus has two boundaries. Consider the set Γ of all open curves γo totally contained in the annulus and whose endpoints lie on the two boundary components. We define the height h of the foliation to be h = inf lρ(γo). The height is essentially the shortest distance, along the Γ annulus, between the two boundary components. We can now show that

Lemma 6.3. A foliation F with a smooth metric and with height h greater than 2π is isometric to a flat cylinder of circumference 2π and height h.

Proof. Denote the saturating geodesics that foliate the annulus F by γf . We first claim that there cannot be another saturating geodesic γi different from the γf geodesics going through the interior of F . Suppose γi is in the same homotopy type as the γf geodesics, but does not coincide with any γ geodesic. Then pick a point p γ such that p F , and consider the f ∈ i ∈ saturating geodesic γf (p) going through p. Since the metric is smooth at p, and γi and γf (p), ⋆ being homotopic, must intersect elsewhere this is in contradiction with Lemma 6.1. If γi is in some other homotopy class it cannot be completely contained in the interior of F (the curves going around the annulus more than once are longer than or equal to 4π). It is also clear that γi cannot intersect any boundary of F more than once, since in that case, there would be γf curves intersecting γi at more than one point (simply pick a γf curve sufficiently close to the boundary in question). Therefore γi must intersect each boundary once and only once. As a consequence a subcurve of γi is totally contained in F and goes from one boundary component to the other. This subcurve, by definition must be longer than or equal to the height h of F . Since we assumed h > 2π, the subcurve is too long and γi cannot be a saturating geodesic. Since there are no other saturating geodesics within F the arguments of Refs. [15,16] apply and imply that the metric is flat, and isometric to that of a finite cylinder, which must therefore be of height h. We now need to establish a slight variation of the amputation theorem. This is a “cutting theorem”, and applies to internal annuli. While most of the proof is analogous to that of amputation in [16] there are some differences in the analysis of curves, so we give the proof below. Our later arguments will only use the statement of the theorem, therefore the proof, which is not so straightforward, may be skipped by the reader only interested in the general arguments. Theorem 6.3. (Cutting Theorem) Consider a Riemann surface with a minimal area metric R ρ and a foliation F of height 2π +2δ with δ > 0. Let be the middle geodesic in F and let 0 C Rc denote the cut surface (or surfaces) obtained by cutting along . Assume the cut surface(s) R C

⋆ The two geodesics cannot be paralell at p since they would have to coincide elsewhere, given that the metric is smooth at p.

64 has a minimal area metric ρ continuous in a neighborhood of the boundaries created by Rc cutting. Then ρ is the restriction of ρ to . 0 Rc Proof. The general setup for the proof is given in Fig.4, where we show the surface with R the foliation F . This foliation defines an annulus, which must be isometric to a flat cylinder of circumference 2π and height h = 2π +2δ (Lemma 6.3). We will use the canonical flat coordinates on this cylinder where the original minimal area metric ρ0 = 1. When the surface is cut along the curve , it may result in two surfaces or in one surface. The proof below does C not distinguish these two possibilities and we will simple denote the cut surface by . Rc † Since every nontrivial closed curve in is actually a nontrivial closed curve in , it Rc R follows that the restriction to of ρ is an admissible metric on . We just have to show Rc 0 Rc it is of minimal area. Assume there is a metric ρ admissible in , continuous near the 1 Rc boundaries produced by the cutting procedure, and with less area than ρ0. Form now the metric ρ = (1 ǫ)ρ + ǫρ , (6.6) ǫ − 0 1 which is admissible due to linearity, and it has lower area than ρ0 by convexity. One has

A(ρ )= A(ρ ) kǫ + k′ǫ2, (6.7) ǫ 0 − where k,k′ > 0 are constants (independent of ǫ).

The idea, as in [16] is to see how close is ρǫ to be an admissible metric on the original surface . If ρ were admissible on we would be done since we would have a contradiction R ǫ R with the fact that ρ is supposed to be the minimal area metric on . 0 R Let l and r denote the boundaries of the foliation F , as indicated in the figure. Consider C C now the metric ρ on . The closed curves that do not touch must be long enough since ǫ R C they are still nontrivial closed curves on . Consider now curves that touch . They may be Rc C completely contained in F or may not be completely contained in F . Consider the second case first; such curves either intersect l, r or both. Suppose γ intersects l only, then γ must C C C travel twice across the foliation Fl and therefore using (6.6) we find

l (γ) (1 ǫ)l (γ) (1 ǫ)(2π +2δ), (6.8) ρǫ ≥ − ρ0 ≥ − and choosing ǫ < π/δ such curves are clearly long enough. If γ intersects both l and r its C C ρ0-length must also exceed (2π + δ) and therefore the same argument applies. We now come to the curves that do not work automatically. These are the curves that cross and are completely contained in the foliation F . Consider such a curve γ, we will show C now how to give a lower bound for its length. Consider now a subfoliation of F of height N

If it were not it would have to bound a disk D in . Since D cannot be fully contained in c (otherwise † R R the curve would be trivial in c) D must either intersect the cutting curve or contain it completely. Both possibilities are impossible,R the first because the original curve did not intersectC , and the second because C is a nontrivial closed curve and therefore cannot be inside a disk. C

65 2ǫ1/4 symmetrically centered around (as shown in part (b) of Fig.4). We define the constant C K via

1+ K = sup ρ1, (6.9) N where ρ1 is evaluated in the canonical coordinates defined by ρ0. The constant K cannot be negative, because otherwise it would mean that ρ < 1 throughout and the ρ could not 1 N 1 be admissible. We will therefore consider only K 0 and bounded (since we assumed the ≥ minimal area metric is continuous in some neighborhood of the boundary). Let us now show that for curves γ completely contained in the foliation F one must have l (γ) 2π(1 K). (Recall that on the surface the metric ρ is possibly discontinuous ρ1 ≥ − R 1 across .) Consider a curve γ, its intersections with divide the curve into segments. Let C C ai and bi denote the pieces of γ on Fl and on Fr respectively. Moreover, let ai and bi denote the segments on homotopic to a and b respectively. We now have two nontrivial closed C i i curves in addition to γ, the first is obtained joining the ai segments to the bi segments, and the second joining the bi segments with the ai segments. Both curves are nontrivial curves on (since they are homotopic to the boundary ), and therefore admissibility of ρ gives the Rc C 1 following conditions [l (a )+ l (b )] 2π, ρ1l i ρ1l i ≥ Xi (6.10) l (b )+ l (a ) 2π, ρ1r i ρ1r i ≥ Xi   where ρ1l and ρ1r denote the metric ρ1 restricted to Fl and Fr respectively. We therefore obtain l (a )+ l (b ) 4π [l (a )+ l (b )] (6.11) ρ1 i ρ1 i ≥ − ρ1r i ρ1l i Xi   Xi and therefore l (γ) 4π max [l (a )+ l (b )] . (6.12) ρ1 ≥ − ρ1r i ρ1l i Xi From the definition of K we have that

lρ r (ai) (1 + K)lρ (ai), 1 ≤ 0 (6.13) l (b ) (1 + K)l (b ), ρ1l i ≤ ρ0 i and therefore

l (a )+ l (b ) (1 + K) l (a )+ l (b ) =(1+ K)2π, (6.14) ρ1r i ρ1l i ≤ ρ0 i ρ0 i Xi  Xi  Back in (6.12) we obtain l (γ) 4π (1+ K)2π =2π(1 K) as we wanted to show. This is ρ1 ≥ − − a powerful result since it tells how small a curve can get by using cleverly the discontinuity of the metric ρ across . Note that in order to have good control, the metric ρ weights in ρ 1 C ǫ 1 weakly.

66 We can now reduce the space of curves that are problematic. Consider the subfoliation ′ of F , which is of height 2ǫ1/3. Note that ′ . We now claim that for ǫ sufficiently N N ⊂ N small all curves crossing and extending beyond ′ are good. The reason is that such C 1/3 N curves extend at least a distance ǫ in the foliation and therefore their ρ0-lengths satisfy l (γ) 2π 1+ ǫ2/3/π2 (Lemma 5.1 of [16]). It is then possible to show (see Eqns. (5.20- ρ0 ≥ 5.23) of [16])p that for sufficiently small ǫ one has lρǫ (γ) > 2π. Therefore the only problematic curves are the ones completely contained in ′. Following [16] one introduces a new metric ′ N ′ ρǫ ρǫ everywhere such that the remaining problematic curves now work; ρǫ is admissible. ≥ ′ χ The area of ρǫ differs from that of ρǫ by terms of order ǫ with χ > 1 and therefore (6.7) ′ implies that for sufficiently small ǫ the area of ρǫ is lower than that of ρ0 (the interested reader should consult [16] for details). This is the contradiction that completes our proof of theorem 6.3.

6.4. Towards a Concrete Description of All String Vertices

We have a very explicit description of the minimal area metrics for genus zero, and of the string vertices for genus zero; these are simply given by the restricted polyhedra [1,2]. The polyhedra represent the critical graphs of a Jenkins-Strebel (JS) quadratic differential. For any surface, with at least one puncture, one can find a unique JS quadratic differential with second order poles at the punctures, all of equal (real and negative) residue, and with a number of ⋆ ring domains equal to the number of punctures of the surface. This simply means that any surface with n punctures can be constructed by gluing together the open boundaries of n- semiinfinite cylinders of circumference 2π in a unique way. The pattern of gluing is encoded on the ‘polyhedron’ of the surface, which is the critical graph of the quadratic differential. Thus any surface has an associated polyhedron. The metric induced on the surface by the quadratic differential (ρ = φ(z) 1/2, where φ(z)dz2 is the quadratic differential) is flat except | | at isolated points with negative curvature singularities. The main result of Refs.[1,2] was that the string vertices are given by the surfaces whose polyhedron is a restricted polyhedron, V0,n that is, a polyhedron such that all of its nontrivial closed paths are larger than or equal to 2π. Such vertices were shown to give a single cover of the moduli spaces relevant to the M0,n classical theory [2,8]. No such concrete description is available yet for higher genus, basically because we do not know completely what type of curvature singularities the higher genus metrics have. A fairly concrete description is possible, however, in terms of the structure of the foliations by geodesics that we expect to exist. The basic result indicating how complicated the pattern of foliations can be was given in Lemma 6.1. An interesting counting problem would be to determine which is the maximum number of foliations that can go through a point in a surface of genus g and with n punctures. This would require counting the maximum number of simple closed Jordan curves of different homotopy all of which have a single common point p and intersect nowhere else. We expect the different patterns of foliations as we move in to correspond Mg,n

⋆ The reader who is not familiar with quadratic differentials may consult Refs. [29,1,8,45].

67 to different “cells” in a decomposition of the moduli space. We will not try to elaborate on this point at present. Instead we claim that a description of the string vertices can be given as follows: The String Vertex . Consider a Riemann surface of genus g 0 and n 0 punctures Vg,n R ≥ ≥ (n 3 for g = 0) equipped with the metric of minimal area. Then, if and only if ≥ R ∈ Vg,n the heights of all internal foliations are less than or equal to 2π. If is in the coordinate R Vg,n curves are defined to be the curves around each puncture. Cπ In the above, internal foliation refers to a foliation that does not correspond to a punctured disk, namely, it is not one of the semiinfinite foliations around the punctures. An internal foliation defines an internal annulus. We have placed the coordinate curves leaving stubs of length π for each puncture. This is necessary in order to make sure that the plumbing procedure produces candidate metrics; if we had no stubs the plumbing of two punctures on a single surface with a propagator of small length could introduce curves shorter than 2π ruining the construction. While the stubs are not necessary for the classical theory, the quantum theory requires them and we must therefore use them here. This actually means that the ’s determined by this rule do not agree with the restricted V0,n polyhedra! This is actually good, for it was clear all along that the restricted polyhedra were problematic at the loop level. We will elaborate on this point in the next subsection. At any rate it is nice that a simple generalization of the restricted polyhedra gives a correct description of all string vertices. Let us verify briefly that this definition for the string vertices satisfies conditions (a),(b), and (c) required in 5.1. Condition (a) is clearly satisfied since the disks determined by § the coordinate curves cannot ever intersect. The string vertices are not of the overlap type. Condition (b), that the assignment of local coordinates be independent of the labelling of the punctures is also manifest; the minimal area metric does not depend on the labels we assign to the punctures, and therefore the coordinate curves do not either. Condition (c), that the local coordinates in mirror pairs be related by the antiholomorphic map between the surfaces, is slightly less trivial. We claim that given a mirror pair and ∗, the minimal area metrics R R on these surfaces must be precisely related by the antiholomorphic map. First note that given an admissible metric on the map gives us an admissible metric on ∗, and viceversa. This R R shows that the minimal area metrics on and ∗ must have the same area. Uniqueness R R of the minimal area metric then implies that the metrics are related by the map. Since the map takes geodesics to geodesics, it must, in particular take the coordinate curves on one surface to the coordinate curves on the other, since these are determined by the last saturating geodesics homotopic to the punctures. Thus the coordinate disks, and as a consequence, the local coordinates are mapped into each other by the antiholomorphic map relating the mirror pair. The case of the one-loop vacuum graph is special. Suppose, as usual, that we build tori using length T cylinders of circumference 2π, and gluing together the boundaries with a twist angle θ. If we do this for all values of T and θ ( π θ < π) such that all closed curves − ≤ on the resulting torus are longer than or equal to 2π, this will produce exactly one copy of

68 the fundamental domain [8]. This happens because τ =(iT + θ)/2π, and the length lc of the closed geodesics along the tube is l = 2π√T 2 + θ2. Then, l 2π implies that τ 1 (the c c ≥ | | ≥ condition on θ gives the standard condition on the real part of τ). In this construction we have a single foliation, and its height is the length T of the cylinder we started with. The definition of the string vertices, applied to would imply that this vertex is not zero; it includes all V1,0 the tori built with cylinders of length T 2π (and having some nonzero twist angle). This ≤ result, from the viewpoint of Feynman rules, would only seem to make sense if we use a cutoff propagator. Using such cutoff propagator is exactly the same as using stubs, except for this vacuum vertex (more on this in 6.5). In the cutoff propagator picture the string vertices are § the same as in our definition above, with the difference that the coordinate curves would be the curves. While we will use the stubbed vertices viewpoint in this paper, we should keep C0 in mind that the alternative viewpoint could turn out to be relevant. Reconstructing Feynman Graphs. Before checking that the above is the correct string vertex let us show how to reconstruct the Feynman graph associated to a surface from the knowledge of its minimal area metric. Given a surface with a minimal area metric first consider the R semiinfinite foliations. They correspond to the external legs of the diagram. Then consider the foliations of height bigger than 2π. If there are none, then the surface is an elementary interaction , where n 0 is the number of infinite height foliations, and g is the genus of Vg,n ≥ the surface. The foliations of height greater than 2π, which must be flat cylinders, give rise to the propagators. On each such cylinder mark two closed geodesics, each a distance π from each boundary of the cylinder. Mark one curve on each semiinfinite foliation, a distance π from its boundary. Cut the string diagram open along all these curves. The surface breaks up into a number of semi-infinite cylinders, the external legs, a number of finite cylinders, the propagators, and a number of surfaces with boundaries, that correspond to the elementary interactions. Each elementary interaction corresponds to an element of the set where g is Vg,n the genus of the surface with boundary and n is the number of boundaries. This defines in a unique way the Feynman graph associated to the surface. In a similar way, a Feynman graph determines a unique minimal area metric; given the data associated to the propagators and the vertices used, we can build the surface. It follows that two different Feynman graphs V must built surfaces with different patterns of foliations, and therefore different metrics. Consistency of the Choice of . Let us now return to the problem of verifying the consis- Vg,n tency of our ansatz for the string vertices. Two things must be checked. First, no surface in must be produced by sewing. This is clear since due to the stubs of length π present in V every vertex, sewing creates foliations of height greater than or equal to 2π. When the height is greater than 2π the surface is clearly not in (by definition), when the height is equal to V 2π we have the case when the boundary of the set coincides with the collapsed propagator V boundary of the sewing domain. The uniqueness of the metric of minimal area plus the com- patibility of sewing with minimal area implies that the metrics agree in this boundary, and so do the coordinate curves, as a consequence. The second point that must be checked is that all surfaces which are not in are produced by sewing, and produced only once. The first part is V clear because for surfaces not in there is at least one foliation whose height is greater than V 2π. By the cutting theorem we can cut all the foliations open and obtain a surface(s) with

69 minimal area metric(s). Restoring the semiinfinite cylinders around all boundaries one obtains the Riemann surface(s) whose plumbing gives us the original surface. Finally, no surface could have been produced more than once, since different Feynman diagrams correspond to differ- ent metrics, which by uniqueness of the minimal area metric, cannot correspond to the same Riemann surface. This concludes our proof that is indeed the string vertex. Since we have established that V the ’s generate a single cover of all relevant moduli spaces, it follows, by the arguments in 5 V § that they satisfy the recursion relations (5.1). It is important to note, for our applications in 9 that our arguments hold also for string § vertices defined by the condition that all internal foliations be of height less than or equal to 2l, where l is any length satisfying l π. In this case the coordinate curves defined by ≥ Cl with l π. This means that the stubs are now of lenght l. We therefore have generated a one ≥ parameter family of consistent string vertices. Thinking of the stub length as a ultraviolet cutoff parameter the basic identity (5.1) can be shown to imply a group equation for string field theory [43,44].

6.5. Rearranging the Classical Theory

Consider the classical theory, that is restrict to the set of vertices , with n 3. The Vg=0,n ≥ consistency condition for such vertices does not involve the second term in the right hand side of the recursion relation (5.1). Given such vertices one can construct a classical field theory that has a ‘BRST’ invariance and a gauge invariance. One can show that under the condition of symmetry (condition (b) in 5.1) there is no solution with only a three-vertex and no higher § vertices [46]. The simplest known solution is given by the restricted polyhedra described in the previous subsection. Let us see what is the problem in trying to extend this to loops. If the classical vertices were sufficient we should be able to see that the remaining consistency conditions are satisfied. Actually, if we could show that all the vertices at genus one vanished, then all the higher V1,n ones would vanish too. Let us look at the simplest vertex, namely . The consistency V1,1 condition requires that ∂ = R (loop) = 0 where R (loop) is the set of surfaces obtained V1,1 − 1 1 sewing two of the legs of the three string vertex. This set of surfaces does not vanish when we use the naive three string vertex. In fact the set of surfaces we get corresponds to noded surfaces! This happens because the naive closed string vertex is an overlap, and is just another manifestation of the fact that the three-vertex together with the standard propagator does not give the correct one loop tadpole [45,47]. The same problem would take place for all the vertices since in the right hand side of the consistency equation we would find a classical V1,n vertex with two legs identified, and since all classical vertices are overlaps this is bound to introduce degenerate surfaces. Let us therefore deform the original vertices in order to obtain sensible ones by using two simple steps discussed in [7] and [5]. This procedure, applied to the original classical vertices, which give problems at loops, gives us the string vertices discussed V in the previous section.

70 Imagine we begin with a set of Feynman rules such that the propagator involves tubes with length l such that l l . Moreover assume the vertices have legs all of which are of length a, ≥ 0 for example, overlap vertices but with cylindrical stubs of length a included. Note that when two vertices are joined by a propagator, the intermediate tube has total length, including stubs, l 2a + l . Suppose we want to change the propagator so that it includes lengths l l l . ≥ 0 ≥ + ≥ 0 If we now join vertices, the intermediate tubes will have lengths l 2a + l , and therefore ≥ + we are going to miss some surfaces. One can compensate by changing the stub length a, one amputates them, by an amount which is half the difference between l+ and l0. Since any intermediate line includes twice the stub length, this will restore the missing surfaces. In the same way, if l0 is changed into l−, which is smaller than l0, the intermediate lines become too short, and this is compensated by increasing the length of every stub by half the difference between l0 and l−. This latter operation is always possible since we can always add stubs to the vertices. The former operation cannot always be done since we cannot remove lengths greater than a to the stubs. In fact, the first step we must do is of the problematic type. We begin with propagators that go from zero to infinity and with contact type vertices (zero length stubs). We want to change the propagator to include only l l > 0. In the above procedure this would involve ≥ 0 amputating stubs, which is not possible. The solution was given in [7]. One leaves the three- point vertex as it is. Then in calculating four point amplitudes we would miss the Feynman graphs involving two three-point vertices and an intermediate tube of length less than l0. We then define these Feynman graphs to be an extra contribution to the original four-point vertex. This procedure can be continued. The general answer given in [7] is that at any order the new vertices are given by the old ones plus all the Feynman graphs built with the lower point original vertices and all intermediate tubes of length less than l0. This is the first step. The second step consists of restoring the original propagator. As discussed above this is never a problem and it is done simply by adding stubs of length l0/2 to every one of the vertices (including the three-point vertex) obtained by step one. In this way we have obtained a new solution of (5.1). If we begin with classical overlap vertices, we obtain classical vertices that are not overlaps. In building up the full solution to higher genus by using minimal area, it is convenient to take l0 = 2π. Thus the rearranged classical vertices will have cylindrical stubs of length π, and we have recovered the result of the previous subsection. The vertex , for example does not only include the restricted tetrahedra (with stubs) but also string V0,4 diagrams built with two (unstubbed) three string vertices and a propagator of length less than 2π joining them.

71 7. Construction of the String Field Interactions

In the present section we use the geometrical structures of 5 and 6 to construct the string § § field products, and the string field vertices, or multilinear functions. The main input is the set of string vertices . Since such vertices correspond to subsets of the moduli spaces Vg,n Mg,n we will have to concern ourselves with integration. As we will see, we want to integrate over a section of a certain bundle over moduli space (the integrand is not defined on moduli space P because we work off-shell). Our first step will be to understand the various spaces involved. b Then we show how a tangent representing a deformation of a surface with local coordinates, can be constructed concretely with a Schiffer variation. Such variations are done with vector fields on the surface, and these vector fields are relevant to define the antighost insertions. We define a family of forms of different degree on , all labeled by a set of state vectors of the P conformal field theory. These forms are proven to be well defined if the state vectors satisfy − − b the subsidiary conditions b0 = L0 = 0. We then show that the BRST operator acts as an on the above forms. We define the string field multilinear functions and multilinear producs, and establish some of their properties. Much of the technology necessary for us has been presented earlier in Refs.[36,37], where closely related issues were studied.

7.1. The Schiffer Variation

Our basic data are Riemann surfaces with punctures and with a canonical set of local coordinates around the punctures. As usual the surfaces are considered as points of , and Mg,n the space of surfaces with local coordinates at the punctures is denoted as . This may be Pg,n called the space of decorated surfaces. Thus a surface with local coordinates, or a decorated surface, represents a point in . The space is infinite dimensional, since the space of Pg,n Pg,n local coordinates at a point is infinite dimensional. We are interested in forming a coset space = / , where indicates that two identical punctured surfaces, with local coordinate Pg,n Pg,n ∼ ∼ systems that differ by a constant phase around each puncture should be identified. The space b is still infinite dimensional. Recall, from the discussion in 5 that local coordinates are Pg,n § defined by “coordinate curves”, Jordan closed curves homotopic to the punctures, with a b marked point on them. The curves correspond to the locus z = 1, and the marked point is | | z = 1, with z the local coordinate. The space corresponds to the space of surfaces with Pg,n these marked coordinate curves. The space is simply the space of surfaces with unmarked Pg,n coordinate curves. b Given a point in there is an obvious projection π′ to , namely the point is mapped Pg,n Pg,n to the coset where it belongs, in other words we forget about the marked point on the coordinate b curves. There is another projection π from to consisting of forgetting about the local Pg,n Mg,n coordinates. Finally there is a projection from to which simply means forgetting b Mg,n Mg,0 about the punctures. In 5 we actually gave a map σ from to since we found a § Mg,n Pg,n way to define canonical local coordinates (up to phases) on any punctured Riemann surface b using minimal area metrics. The local coordinates are defined in terms of the maximal domain foliated by geodesics homotopic to the punctures. We summarize the above information in the following diagram

72 Pg,n π′   Pyg,n σb π (7.1) x    Mg,ny   Myg,0 ⋆ The map σ, defines a canonical section in . Clearly (π σ) is the identity map on the Pg,n ◦ space . Mg,n b We must understand the various tangent spaces appearing here. Consider a point P ∈ corresponding to a decorated surface. Let P = π′(P ) denote the corresponding Pg,n ∈ Pg,n equivalence class, P = π(P ) denote the underlying punctured surface and P~ = f(P ) ∈ Mg,n b b ∈ denote the underlying surface. Let us introduce the following notation: Mg,0 e b e V ,V , T ( ), 1 2 ···∈ P Pg,n V1, V2, T ( g,n), ···∈ P P (7.2) Vb1, Vb2, T ( b g,n), ···∈ Pb M ~ ~ Ve1, Ve2, T ~ ( g,0). ···∈ Pe M

The tangent vectors in T ( ) correspond to changes in moduli of the unpunctured P~ Mg,0 surface P~ . The tangent vectors in T ( ) include the deformations that change the positions P Mg,n of the punctures and the underlying moduli of the punctured surface P . The tangent vectors in e T ( ) include the possibility of changing the unmarked coordinate curves (changing the local P Pg,n e coordinates by more than a phase) in addition to changing the positions of the punctures and b b the underlying moduli of P . Finally the tangent vectors in T ( ) correspond to all possible P Pg,n deformations of the marked coordinate curves, changes in the positions of the punctures and b changes in the underlying moduli of P . The projections between the various spaces in (7.1) induce obvious projections between the respective tangent spaces. The projections can map many vectors to the same one, due to the existence of “vertical” vectors, which are those that project to zero. They are the following; in T ( ), the vectors that only change the marked points on the coordinate curves; in T ( ) Pg,n Pg,n the vectors that only change the unmarked coordinate curves; and in T ( ) the vectors that Mg,n b

′ ⋆ Technically speaking g,n is a principal fiber bundle with base space g,n, projection (π π ), and with fiber F the possibleP local coordinates, which is just the topologicalM group G of maps of◦ the form n z anz with (a1 = 0). → n=1 6 P 73 only change the positions of the punctures. Such vectors are the ones that change precisely the data that is forgotten by the projection maps. Our concrete objective is to understand how given a decorated surface P we can im- ∈ Pg,n plement concretely on the surface the variations that correspond to the tangent space T ( ). P Pg,n The appropriate tool is the Schiffer variation [48,49,50], which uses vector fields on the surface to generate the changes in data. The main ideas have been summarized in the physics litera- ture in [36,37]. Let P correspond to the punctured surface , with punctures p ,p , p and R 1 2 ··· n coordinate curves γ ,γ γ . We are now given an n-tuple of vector fields 1 2 ··· n v =(v(1), v(2), v(n)), (7.3) ··· where the vector v(i) is a holomorphic vector defined (a priori only) in an annular neighborhood of the coordinate curve γi surrounding the puncture pi and defining the disk Di (some of the vectors v(i) may be zero). One defines, for each of the punctures the maps

z z′ = z + ǫv(i)(z ) (7.4) i → i i i which for sufficiently small ǫ map the coordinate curves γi into simple closed Jordan curves γǫ. The new surface ′ is constructed by gluing the interior of γǫ (the domains which include i R i the respective punctures) to the exterior of γi, for each puncture:

′ = Int(γǫ) D . (7.5) R i ∪ R − i  Xi  Xi

This gluing is done by identifying points as follows. Let p be a point on the boundary of D , and p′ be a point on the boundary of the new disk defined by γǫ. We identify the R − i i two pointsP p γ p′ γǫ when z′(p)= z (p′). (7.6) ∈ i ←→ ∈ i i i More explicitly, we remove the disk defined by the coordinate curve, namely the zi = 1 disk ′ ′ | | Di and map its boundary into a curve in the zi plane defining a new disk Di. It is crucial to note that while we are only mapping the boundary, this defines a new disk. Any two punctured disks are conformally equivalent, but here we have two disks whose boundaries are mapped into each other in a prescribed analytic way. Under this condition it is not clear that this map extends to the complete disks. We then glue the disks D′ back into D . This gives us i R − i the new surface. Intuitively it is very clear what the variation is doing,P it leaves Di R − unchanged but is sticking new disks back (this is the reason for the term “Schiffer’s interiorP variation”). We now need to understand what happens to the punctures and the local coordinates. ′ The new local coordinates are simply the zi coordinates. This is a well defined coordinate in ′ ′ the new disk Di. The new puncture pi is located at the origin of the new , ′ (i) namely at zi = 0. Since the vectors v do not necessarily extend holomorphically inside the

74 disk, there is no other sensible alternative for the new coordinates or new puncture. It should ′ be emphasized that the new canonical coordinates are simply the zi coordinates and therefore ′ the boundary of the disk Di does not coincide with the new coordinate curves. Those are now the z′ = 1 curves. An alternative way to get the new coordinate curves directly would be | i| to do the Schiffer variation with curves γ chosen such that under the variation z′(γ ) = 1. i | i i | Then the new coordinate curves are simply the γi curves, which correspond to the boundaries b b of the amputated surface Di. R − b The curve we use to do theP Schiffer variation can be changed. We can begin with another curve homotopic to the original curve and the resulting surface is identical as long as one curve can be deformed into the other one without encountering singularities of the vector field. Suppose we have a curve γ defining a punctured disk D with puncture p and a vector field v that has a singularity inside D at some point q = p and possibly a singularity at p. Then by 6 contour deformation the original variation using γ is equivalent to doing two variations, one with a contour enclosing q and not p, and another with a contour enclosing p but not q. This means that the most general situation we need to consider is really that of a Schiffer variation with a vector field that inside the countour may become singular only at the puncture.

7.2. Using the Schiffer Variation to Generate a Tangent

Let us begin by discussing the conditions under which data do not change. Suppose all ′ (i) the mappings zi zi extend holomorphically to the interior of the disks, namely all v are → ′ holomorphic in Di. Then the disks Di and Di are completely equivalent and when reinserted back, if we forget the punctures, the surface has not changed. Thus such vectors do not change the moduli of the underlying unpunctured surface. Suppose, in addition that a vector (k) v vanishes at zk = 0, then the position of the puncture pk is not changed. This is clear ′ because in this case the disks Dk and Dk are completely equivalent as punctured disks since ′ zk = 0 is mapped to zk = 0 and the puncture has not moved. (i) Suppose a vector v extends holomorphically into the disk Di and does not vanish at ′ the origin zi = 0. Then the conformal map relating the Di and Di disks maps the original puncture into z′ = ǫv(i)(z = 0) = 0. Since the new puncture is now at z′ = 0 this indicates i i 6 i the possibility that the puncture may have moved. We will see presently how to decide if it moved or not. If all the vectors v(i) comprising v arise from a single globally defined holomorphic vector field on the whole of p (singularities are permitted at the punctures) then the variation, R− i performed around all theP punctures changes nothing at all! It does not change the underlying moduli, nor the position of the punctures, nor the coordinate curves, nor their marked point. Let us derive this. Consider Fig.5 where we show a surface with one puncture, for simplicity. The puncture is surrounded by the curve γ and we will do a variation using a vector field v that extends holomorphically to the exterior of γ, but not necessarily to the interior of γ, where it may have singularities. We will show that this variation does not change any data. In part (a) we show the original surface decomposed into the disk D, and D, to the right R − (both shaded). In part (b) we have the new surface ′ obtained by the variation, decomposed R

75 into the disk D′ and D. The respective coordinates are denoted as z′ and z′ and the R − gluing is done by identifying points s′ ∂D′ with s′ ∂( D) via z′(s′)= z′(s′)+ǫv(z′(s′)). ∈ ∈ R− Let us now show how to construct a conformal map between the punctured surfacese and R ′. First we rearrange the construction of givene in (a) into what is showne ine (c); wee havee R R changed the shape of the initial disk D into that of D′ by adding and subtracting regions and compensating for this on D (this is just the same as cutting the original surface along R − R the image of the original curve γ under the variation.). The coordinate systems for the disk and the remainder are called z and z, and the gluing of points s and s in their boundaries is via z(s)= z(s). Now the mapping between ′ and is very simple, we use Fig.5 (b) and (c). R R For the points on the disks we simplye take the identity map e e e q′ q when z(q)= z′(q′) (7.7) → Note that this map maps the new puncture into the old puncture and furthermore the coor- dinates of points mapped into each other are identical! If we show this map can be extended to the complete surfaces we will have shown that the puncture has not moved and that the coordinates around it have not changed. For the remainder of the surfaces we take

q′ q when z(q)= z′(q′)+ ǫv(z′(q′)). (7.8) → This map exists because thee vectore field wase assumede e e to extende toe the remainder of the surface. It is simple to check that the complete map from ′ to defined by the last two equations R R respects the identifications along the cutting lines, and is therefore a well defined map of the surfaces. This concludes the argument. It is also clear how to extend the argument for the case when there are several punctures and the vectors around the punctures arise from a single holomorphic vector field on p . R − i Actually, the complete resultP is that an n-tuple of vectors defined around the punctures does not change any data if and only if it arises from a globally defined holomorphic vector field on the surface minus the punctures. We showed above that such vectors do not change the data, now we argue that any vector which is not of this type must actually change the data. Consider a vector field defined around a puncture, assume it extends holomorphically up to the puncture, where it vanishes, but does not extend holomorphically to the rest of the surface. This vector does not change the underlying moduli, nor the position of the puncture, but it does change the local coordinate around the puncture. The map between the original surface and the new surface is simply the map between the two disks, induced by the vector field, plus the identity map on the leftover of the surfaces. It is clear that that the coordinates have changed, unless there are continuous conformal self-mappings of the whole surface into itself. Such mappings only exist for g = 0, n = 1, 2, 3 and g = 1, n = 0, and are generated by globally defined vector fields (conformal Killing vectors). However, since these maps would restore back the coordinates they should agree with the vector that generated the variation around the puncture, in contradiction with the fact that the variation vector does not extend.

76 A similar argument would show that a vector that extends holomorphically into the puncture without vanishing there, and does not extend to the rest of the surface must change the position of the puncture, unless the surface is one of the above exceptional cases, in which case at least the local coordinates are changed. Consider changing the underlying moduli of the surface. It is possible to show that the 3g 3 complex moduli of the underlying surface can be changed by variations around a − single puncture [48]. That is, there are vector fields around a puncture which do not extend holomorphically to the rest of the surface nor to the puncture, that can be used to represent the 3g 3 (complex) dimensional vector space T ( ). The Weierstrass gap theorem shows − Mg,0 that if we specify the leading singularity z−n, with n > 0, of a vector field around a point p (with z a local coordinate) there are 3g 3 values for n such that the vector cannot be − ⋆ extended holomorphically to the rest of the surface. These are the vectors that will produce the underlying change of moduli of the surface. Every vector v = z−n, with n one of the special values, must generate a nontrivial tangent in T ( ). The tangents associated with Mg,0 all the special values of n must give a basis for T ( ). If this were not the case, we would Mg,0 not be able to generate all possible tangents by Schiffer variations. This shows that any vector that does not extend (in and out) must generate a deformation of the underlying moduli. We summarize all the above information in the following table. The first two columns tell us whether the Schiffer vectors extend holomorphically into the punctures (extends in) and/or to the rest of the surface (extends out); v(p) denotes the value of the Schiffer vector at the puncture. To the right we show the type of data they change. Table. Schiffer variations.

Changes Moves Changes Extends in Extends out moduli punctures coordinates Yes/No Yes No No No

Yes, v(p)=0 No No No Yes

Yes, v(p) =0 No No Yes∗ 6 −− No No Yes∗∗ −− −−

The asterisk on the third row indicates that the positions of the punctures change unless n = 1, 2, 3 with g = 0, or n = g = 1, in which case the local coordinates must change. The double asterisk indicates that the underlying moduli change unless the surface is of genus zero

⋆ For genus one, a vector field whose leading singularity is z−1 cannot be extended. If it could, then working on the presentation of the torus as a lattice in the z-plane the vector field would be of the form v(z)(dz)−1 where actually v(z) must be a function on the torus. But it is well-known that there are no functions on a torus having a single first order pole (the famous function of Weierstrass has a double pole). For genus zero, all vectors with leading singularity z−n withP n> 0 can be extended.

77 and there are no such moduli. If this is the case the vector must change the position of the punctures, unless n =1, 2, 3, in which case the local coordinates must change. We can explain now how to find the variations and the associated vectors on the surface corresponding to a given tangent vector in . We use a given puncture to generate the Pg,n underlying moduli deformations. We then move the positions of the punctures by a Schiffer variation around each puncture with a vector field that extends holomorphically inside the disk but does not extend outside. If that vector does not vanish at the puncture it will change the position of the puncture. We finally change the coordinate system around the punctures. This is also done with a variation around each puncture using a vector that extends holomorphically inside the disk, vanishes at the puncture and does not extend outside. Two successive infinitesimal Schiffer variations with vectors ǫv and ǫv′, are equivalent, to order ǫ, to a single variation with vector ǫ(v + v′). Since we are interested in reproducing a specified tangent in , all variations are infinitesimal, and therefore the n variations moving the Pg,n punctures and the n variations modifying the coordinate systems can be put together into n variations, each around a different puncture. Once we have learned how to do the variations there is no sense of order in which the underlying moduli variations and the puncture variations must be done. To summarize, consider a specified tangent vector in . It can be generated by the sum Pg,n of (3g 3) variations of the type −

(v(1), 0, 0, , 0), k =1, 2, , 3g 3 (7.9) k ··· ··· − all based on the same puncture (and each changing a complex modulus) , and n variations of the type (0, 0, , v(k), , 0), k =1, 2, , n (7.10) ··· ··· ··· each around a different puncture (changing the position of the puncture and the local coordi- nates). Thus a given tangent in maps into the vector v which is simply the sum of the Pg,n 3g 3+ n vectors listed above. Any other possible vector v′ representing the same tangent in − can only differ from v by a vector t arising from a globally defined holomophic vector on Pg,n the surface minus the punctures.

7.3. Construction and Properties of Forms on Pg,n b Our aim in this section is to construct a set of differential forms on the space . The Pg,n degree of these forms will go from zero to infinity, since the space is infinite dimensional. Pg,n b The most relevant forms, however, are those of degree less or equal to 6g 6+2n, the b − dimensionality of the moduli space of surfaces of the corresponding genus and number of punctures. These forms can be naturally integrated over the minimal area section, or subspaces of this section. The form of degree 6g 6+2n is the basic ingredient in the definition of string − field vertices and string field products, as we will see in the next subsection. The forms will

78 be labeled by n states. These will be denoted as B B B B~ ⊗n. (7.11) | 1i⊗| 2i⊗···⊗| ni≡| i ∈ H Here B denotes the state to be inserted on the k-th puncture. | ki Basic Objects in the Operator Formalism. We must recall some basic properties and construc- tions of the operator formalism which are necessary for our purposes. Consider our Schiffer variations, in a surface with one puncture, for simplicity. We write (1 + ǫδ )z z + ǫv(z), (7.12) v ≡ where this is the way the variation acts on a local coordinate. It follows by a simple compu- tation that these variations satisfy the algebra

δv1 , δv2 = δ[v2,v1], (7.13)   as verified by acting on a local coordinate. For the case of several punctures, the vectors become the boldface vectors vi, and (7.12) is naturally generalized by acting on an n-tuple of local coordinates. We then have that

v v δ 1 , δ 2 = δ[v2,v1]. (7.14)   Associated to a point P in , which represents a surface Σ of genus g with n punctures (here Pg,n n 1) with local coordinates, the operator formalism gives us a state Σ ⊗n (defined up ≥ | i ∈ H to normalization). The corresponding conjugate state is obtained by use of the reflector

Σ = R ′ R ′ Σ ′ ′ . (7.15) 1···nh | h 11 |···h nn | i1 ···n We must understand how the state Σ varies as we change the data of the surface, in particular, | i if we make a Schiffer variation with a vector v =(v(1), v(n)). To this end we must use the ··· stress tensor (T (z), T (z)) of the theory. Let us define n (i) (i) dzi (i) (i) dzi T(v)= T (zi)v (zi) + T (zi)v (zi) , (7.16)  I 2πi I 2πi Xi=1

(i) (i) where the operator valued fields T (zi) and T (zi) refer to the i-th Hilbert space (we use dz/2πiz = dz/2πiz = 1). The vectors v(i) are simply the complex conjugates of v(i) and Hthe integralsH are taken using contours surrounding the punctures and lying in the domain of definition of the local coordinates and the n-tuple of vectors. We then have that T(v ), T(v ) = T v , v , (7.17) 1 2 − 1 2     which follows from the operator product expansion 2 1 T (z)T (w) T (w)+ ∂T (w)+ , (7.18) ∼ (z w)2 (z w) ··· − − the analogous relation for the antiholomorphic modes, and the fact that the various Hilbert spaces in (7.16) are independent. It is convenient at this point to introduce the antighost

79 analog of (7.16). We define b(v) as

n (i) (i) dzi (i) (i) dzi b(v)= b (zi) v (zi) + b (zi)v (zi) . (7.19)  I 2πi I 2πi Xi=1

Making use of the operator product expansion

2 1 T (z)b(w) b(w)+ ∂b(w)+ , (7.20) ∼ (z w)2 (z w) ··· − − we find that

T(v), b(vk) = b vk, v . (7.21)     Similarly, making use of Eqn.(2.23) we readily obtain

n Q(i), b(v) = T(v). (7.22) { } Xi=1

Returning now to the stress tensor, consider its BPZ conjugation property. Since the stress tensor behaves on a Riemann surface like a quadratic differential (a dimension two primary field) its behavior on the reflector must be identical as that of the antighost field ((2.58)), we therefore have (1) (2) R (Ln L )=0. (7.23) h 12| − −n Using the mode expansion for the stress tensor and that for an arbitrary vector

L v T (z)= n ; v(z)= n , (7.24) zn+2 zn−1 Xn Xn we find that dz T (v)= T (z)v(z) = L v . (7.25) 2πi n −n Z Xn Using (7.23) and (7.25) we find that

R T (1)(v)+ T (2)(vT ) =0, (7.26) h 12|   where the vector vT (z) must be given by

v 1 vT (z)= −n = z2v . (7.27) − zn−1 − z Xn 

80 It is a straightforward calculation, using the above definition of vT in terms of v to find that

T T T v1 (z), v2 (z) = v1(z), v2(z) . (7.28)     We can now address the issue of deforming the state Σ . Following [37,36] one must have | i T δv Σ = T(v ) Σ . (7.29) | i | i One can make a simple consistency check

T T δv , δv Σ = T(v ), T(v ) Σ 1 2 | i 1 2 | i   = T vT , vT Σ 2 1 | i T (7.30) = Tv , v  Σ 2 1 | i = δ vv Σ ,  [ 2, 1]| i

in agreement with the algebra of the δv operators in (7.14). In fact, the other candidate: δ Σ = T (v) Σ would have also passed this consistency check. We can see that (7.29) must v| i | i be the correct form by a consistency argument. Consider a surface with one puncture, with local coordinate z giving a state Σ . The one point function of a physical state Ψ is essentially | i | i given by Ψ Σ . Suppose we change z by the addition of a vector field v(z) = ǫzn+1 h |···| i (n 1). In this case T (vT )= ǫL . If we change the coordinates in this way, the coupling ≥ − −n to a physical state Ψ must not change. This implies that Ψ L Σ must be zero, and h | h |··· −n| i this indeed can happen because for physical states we have L Ψ = 0. n≥1| i It follows from (7.15),(7.29), and (7.26) that

δv Σ R ′ R ′ δv Σ ′ ′ 1···nh |≡h 11 |···h nn | | i1 ···n ′ T = R11′ Rnn′ T (v ) Σ 1′···n′ h |···h | | i (7.31) = R ′ R ′ T(v) Σ ′ ′ −h 11 |···h nn | | i1 ···n = Σ T(v), −1···nh | and therefore we have that δv Σ = Σ T(v). (7.32) h | −h | This equation will be useful to us shortly. Let us conclude this brief discussion of the operator formalism with two useful properties. For t a Borel vector (a vector whose data along the punctures extends holomorphically to the rest of the surface) we have that

Σ T(t)=0, Σ b(t)=0, (7.33) h | h | as explained in [36]. Moreover the state representing the surface is annihilated by the BRST

81 operator Σ (Q + Q )=0. (7.34) h | 1 ··· n From the operator formalism we know that the state Σ is a state with (first quantized) | P i ghost number 6g 6+6n − Σ G Σ = (6g 6+6n) Σ , (7.35) ∈ Pg,n → | i − | i and therefore, in our conventions, the bra Σ has the same ghost number. Since this is an h | even ghost number, the bra Σ is Grassmann even. h | It is important for us to have the normalization of the state Σ unambiguously determined. h | We can fix by hand the normalization of the surface corresponding to the three punctured sphere (the three-string vertex) by defining the string coupling constant κ. The normalization of the sewing states was fixed in 2. Then using sewing and the three punctured sphere we can § define the normalization of states for higher number of punctures and higher genus since it is clear that we can produce some surfaces for each (not all because the vertex is fixed), Mg,n and then using the transport equation (7.29) we can extend the definition of normalization to the remaining surfaces in the moduli space (see [36] and [51] for further analysis of the consistency of sewing).

Defining the Differential Forms Ω . We will define now a differential form Ω(0)g,n labelled by B~ B~ n states B , of real degree (6g 6+2n) on the tangent space T ( ) based at the point P i − P Pg,n (recall (7.1)) by the following expression b b b (0)g,n Ω (V1, V6g−6+2n) Ng,n ΣP b(v1) b(v6g−6+2n) B~ , (7.36) B~ ··· ≡ h | ··· | i b b where the necessity for the normalization constant Ng,n

−dg,n −(3g−3+n) Ng,n = 2πi = 2πi , (7.37)   first pointed out by Kugo and Suehiro [3] in the context of the classical action, will become ⋆ apparent in discussing sewing and the main identity. Moreover, if this constant were not present, the string field vertices would not be real (see 7.4). Similarly we can define other § forms of lower degree, all of them labeled by n states:

(−r)g,n Ω (V1, V6g−6+2n−r)= Ng,n Σ b(v1) b(v6g−6+2n−r) B~ . (7.38) B~ ··· h P | ··· | i b b This is a form of degree (6g 6+2n r), it can be therefore used to integrate over submanifolds − − of codimension r in the section σ( ) (thus the label ( r)). Dg,n −

− ⋆ The apparent disagreement with the constant given in [3] is due to our different convention for b0 and different definition for the string field.

82 Given that the state Σ has ghost number 6g 6+6n, the 6g 6+2n antighost insertions h P | − − in Ω(0)g,n reduce the ghost number to 4n. In order to get a nonzero answer we need a total B~ ghost number of 6n, since in between two vacua 1 1 for each Hilbert space there must h |···| i be a total ghost number of six. Thus for the form Ω(0)g,n to be nonvanishing the sum of the B~ (−r) ghost numbers of the B’s must be 2n, that is (G(Bi) 2) = 0. The form Ω can be i − B~ nonvanishing only if n G =2n r. P i=1 i − The bra Σ , livesP in ⊗n, where each Hilbert space corresponds to each of the labeled h P | H punctures. We can label the Hilbert spaces more explicitly as Σ (1, 2, n) . We may also h P ··· | add to the states the label of the Hilbert space where they are inserted, say B , indicates | i(k) that the state B will be associated to the k-th puncture. Operationally, the tensor product in (7.11) means that in calculating an overlap we use

B~ = B B B , (7.39) | i | 1i(1)| 2i(2) ···| ni(n) and therefore equation (7.36), for example, is evaluated as

(0)g,n Ω (V1, V6g−6+2n)= Ng,n Σ (1, n) b(v1) b(v6g−6+2n) B1 Bn . (7.40) B~ ··· h P ··· | ··· | i(1) ···| i(n) b b Let us explain the various ingredients entering into the definition of the ΩB~ forms, noting as we go along the conditions that must hold so that it they are well defined forms on T ( ). P Pg,n The forms Ω are labelled by the n states B that are introduced at the punctures. Given | ii b b a set of (6g 6+2n) real vectors V T ( ), the form Ω must give us a number. One − k ∈ P Pg,n must first choose a point P such that it projects down to P under the map π′ in (7.1). ∈ Pg,n b b b (Must check that: (i) the choice of P does not affect the result.)b Given a vector Vk one must now find a vector V T ( ) that projects via π′ into V . (Must show that: (ii) that the k ∈ P Pg,n k b choice of V does not affect the result.) Then we must choose an n-tuple v =(v(1), , v(n)) k b k k ··· k that represents the tangent Vk as a Schiffer variation. (Must check that: (iii) the choice of vk does not affect the result.) Let us show that (i), (ii) and (iii) hold. Let us begin with (iii); that is showing that the ambiguity in the choice of the n-tuple vk that represents the tangent Vk does not matter. Indeed the only ambiguity, as we discussed in 7.2, is the addition of an n-tuple ‘t’ arising from § a globally defined vector field holomorphic except at the punctures. This ambiguity requires that the replacement of b(vk) by (b(vk)+ b(t)) should not alter the value of the form Ω. Indeed, since Σ b(t) = 0 ((7.33)), the extra term can be brought all the way to the bra Σ h P | h | and gives no contribution.

Let us now turn to (ii). Given a vector Vk the vector Vk is determined up to a vertical vector V , which as explained in the previous subsection, can only change the marked point in the b coordinate curve. This means that the vector only changes coordinate systems up to a phase. Since we can change the phases of the local coordinates around each puncture independently,

83 the general situation corresponds to changing the phase around one of the punctures. The associated vector v = (0, , v(i), , 0) for the Schiffer variation must implement the variation ··· ··· z′ = exp(iǫ)z 1+ iǫz , v(i) = iz . (7.41) i i ∼ i → i

Thus changing Vk to (Vk + ǫV ) changes vk into (vk + ǫv) and as a consequence will change the relevant antighost insertion in ΩB~ as follows

b(v ) b(v )+ ǫ b(v). (7.42) k → k (i) Equations (7.19) and (7.41) give us b(v)= i(b(i) b ), and therefore the extra term in (7.42) 0 − 0 will not affect the evaluation of ΩB~ if

(i) (b(i) b ) B = b−(i) B =0, (7.43) 0 − 0 | ii 0 | ii which must hold for all i. This is the standard condition that states should be annihilated by − b0 , which we must require (this completes the verification of (ii)). Let us finally verify that (i) gives no problems. As we change P into P ′ (both projecting to P ) there are two types of changes involved in (7.36) : the bra ΣP is changed into ΣP ′ , h | ′ h | and, corresponding to each of the vectors V we must now find new vectors V T ′ ( ) (the b k k ∈ P Pg,n old vectors V are in T ( )) and the new n-tuples v′ corresponding to the new tangents. k P Pg,n b k For any two points P and P ′ related by an infinitesimal tangent ǫV T ( ) the variation ∈ P Pg,n of the Σ is given by h P | Σ ′ = Σ ǫ Σ T(v), (7.44) h P | h P | − h P | where v ((7.41)) is the Schiffer n-tuple representing the tangent V . In order to understand the second type of change, consider the original Schiffer variation vk. What we need now is a new ′ variation v k that implements the same change in the curves used for the variation, despite the fact that the coordinates have changed. The answer is

′ v k = vk + ǫ [vk, v]. (7.45)

′ This is derived beginning with the fact that if point p is mapped into p (near a puncture pi) we want this to happen both with the initial variation and with the new variation:

(i) ′ zi(p)+ ǫv (zi(p)) = zi(p ), k (7.46) ′ ′(i) ′ ′ ′ z i(p)+ ǫv k (z i(p)) = z i(p ),

′ (i) where the new and old coordinates are related by z i(p)= zi(p)+ ǫv (zi(p)). A short calcu- ′(i) (i) (i) (i) lation shows that v k = vk + ǫ[vk , v ], which is the content of (7.45) around the puncture

84 pi. As a consequence each antighost insertion changes into b(v ) b(v )+ ǫb([v , v]). (7.47) k → k k Equations (7.44) and (7.47) tell us how to vary (7.36). The strategy is to move the T terms all the way towards the states. In commuting T and b’s we use (7.21). The extra variations (7.47) and the extra terms arising from the commutators cancel. At the end we simply have T acting on the states T(v) B~ . But given the explicit form of v in (7.41), (supported only | i on the i-th puncture) we have that

(i) T(v)= i(L(i) L ), (7.48) 0 − 0 which simply means that the form Ω is well defined if the states in B~ satisfy the usual B~ | i L L = 0 condition. This concludes our proof that the form Ω is well defined on T ( ). 0 − 0 B~ P Pg,n (−r) BRST Action on the Forms Ω . We have introduced above a set of forms Ω ofb degreeb B~ B~ deg[Ω(−r)] = Dim [ ] r, where r Dim . Since is infinite dimensional, there is Mg,n − ≤ Mg,n P no top form, and r can take all possible negative values. Our minimal area section σ in is b Pg,n of dimension equal to that of , and therefore r = 0 gives us top forms for integration over Mg,n b the section. For r 0 the forms can be integrated over some relevant codimension r subspace ≥ (−r) n of the section. The form Ω is nonvanishing only if G(Bi)=2n r. The forms with B~ i=1 − r = 1 are expected to be relevant to string field redefinitions.P − The result that we want to establish is that the BRST operator acts in a simple way on the above forms. We will show that

(∗) ∗ (∗−1) Ω ~ =( ) d Ω ~ , (7.49) Qi|Bi − |Bi P where ‘d’ is the exterior derivative in , Q is the closed string BRST operator, and ‘ ’ is Pg,n ∗ just a number. This relation holds for all relevant values of g and n, and we have therefore b ommitted those labels on the forms. In order to prove this (7.49) we need to establish some preliminary results. Let us first recall that the exterior derivative of any (k 1)-form Ω on some manifold can be defined with ⋆− the help of k vector fields on the manifold

k dΩ(V , , V )= ( )i+1V Ω(V , , V ◦, , V ) 1 ··· k − i 1 ··· i ··· k Xi=1 b b b b b b (7.50) + ( )i+j ([V , V ], V , , V ◦, , V ◦, , V ). − i j 1 ··· i ··· j ··· k 1≤Xi

85 superscript ‘ ’ on a vector field indicates that the vector field is ommitted. Therefore, while ◦ the form dΩ has k entries, the form Ω appearing in the right hand side shows (k 1) entries, − as it should. In applying (7.50) we will be working on the space . Pick a reference point Σ Pg,n ∈ Pg,n corresponding to a surface with coordinate curves. We consider a set of vector fields V , V b 1 ··· k defined on the tangent spaces corresponding to points in a neighborhood of Σ. We introduce b b for convenience a set of coordinates uI around the point Σ. Then a tangent vector at Σ will be denoted as ∂ V V I , (7.51) i Σ ≡ i ∂uI Σ b where the sum over I is implicit, ∂/∂uI is the usual set of basis vectors, and V I are the |Σ i components of the vector. When we want to emphasize that this is a vector field we write

I I ′ ∂ Vi Σ′ Vi (u (Σ )) I , (7.52) ≡ ∂u ′ Σ b where Σ′ is any point in the neighborhood of Σ. Associated with these vectors there are Schiffer variations. We define ∂ v v , (7.53) I |Σ ≡ ∂uI  Σ as the Schiffer vector that implements the deformations associated to the indicated basis vector based at point Σ. By linearity, we have that

v V )= V Iv v , (7.54) i|Σ i I |Σ ≡ i|Σ where the last relation has introducedb a further piece of notation. There is a nontrivial relation between the Schiffer vectors and the tangent vectors due to the fact that the algebra of Schiffer variations must give a representation of the algebra of tangent vectors. We must have that

δ , δ = δ , (7.55) v(Vi) v(Vj ) v([Vj ,Vi])   with the particular ordering of tangentb vectorsb on theb b right hand side arising because the algebra of transformations shows a reversal of ordering with respect to the algebra of the underlying vectors (as in (7.14)). Let us evaluate the left hand side of Eqn.(7.55) acting on a local coordinate z. We begin by calculating v (1 + ǫδv )(1 + ǫδv ) z = (1+ ǫδv ) z + ǫ j Σ (z) (Vi) (Vj ) (Vi) |  (7.56) = z + ǫ vj (z)+ ǫ vi z + ǫ vj (z) , b b |Σb |Σj |Σ  where Σ denotes the point obtained from Σ after the variation corresponding to the j ∈ Pg,n tangent ǫVj. Note that in the second variation we have used the Schiffer vector corresponding b 86 to the new basepoint. By further expansion we get v v 2 v v 3 (1 + ǫδv )(1 + ǫδv ) z = z + ǫ j Σ (z)+ ǫ i Σ (z)+ ǫ j∂ i + (ǫ ), (7.57) (Vi) (Vj ) | | j O where in the lastb term ∂ ∂/∂zb , and the Schiffer vectors in that term, to the approximation ≡ we are working, can be taken to be based at Σ. Exchanging i and j in the above, subtracting the resulting equation from (7.57), and substituting back in (7.55), we find

2 (vj vj )+(vi vi )= ǫ vi, vj + ǫ v [Vj, Vi] + (ǫ ), (7.58) |Σ − |Σi |Σj − |Σ O    where in first bracket in the right hand side the Schiffer vectorsb areb operators on the surface. This relation will be used shortly. We can now begin in earnest our proof of Eqn.(7.49). Consider the case when the form in the left hand side is a k-form, so in particular, = k Dim . Moreover, since (7.49) is ∗ − Mg,n homogeneous and does not involve changes in g or n, the normalization factor Ng,n is irrelevant and we will not show it below. Our strategy will be as follows. We will begin with the form

Ω(V , , V )= Σ b v(V ) b v(V ) B~ , (7.59) 1 ··· k−1 h | 1 ··· k−1 | i   where, for clarity, we haveb writtenb the Schiffer vectorsb as functionb of the tangent vectors. Then we will use Eqn.(7.50) to evaluate dΩ. Finally we will evaluate, by direct computation the left hand side of (7.49). Consider therefore the first term in the right hand side of (7.50) which corresponds to

k ( )i+1V Σ b v(V ) b◦ v(V ) b v(V ) B~ . (7.60) − i h | 1 ··· i ··· k | i Xi=1     b b b b In order to take this directional derivative it should be emphasized that not only the state changes, but also the antighost insertions change, since they depend on vector fields. The derivative can be evaluated as the limit k i+1 1 ◦ ( ) lim Σi b v(V1) b v(Vi) b v(Vk) B~ − ǫ→0 ǫ  h | ··· ··· | i Xi=1    Σi b b b (7.61)

Σ b v(V ) b◦ v(V ) b v(V ) B~ , −h | 1 ··· i ··· k | i     Σ b b b where Σ , as defined earlier, denotes the point in obtained by varying Σ with the tangent i Pg,n ǫVi. We know that Σ = Σ 1 ǫT(v(V )) + (ǫ2), (7.62) b h i| h | − i O  and, in analogy, we define for the antighosts b

2 b(v(Vj)) b(v(Vj)) + ǫ ∆ibj + (ǫ ). (7.63) Σi ≡ Σ O

With a short computation web find that (7.61);b which is simply the first term in the right hand

87 side of (7.50), can be written as

k ( )i Σ T(v(V )) b v(V ) b◦ v(V ) b v(V ) B~ − h | i 1 ··· i ··· k | i Xi=1    b b b b (7.64) + ( )i+j Σ (∆ b ∆ b ) b v(V ) b◦ v(V ) b◦ v(V ) b v(V ) B~ . − h | j i − i j 1 ·· i ·· j ·· k | i 1≤Xi

( )i+j Σ b v( V , V ) b v(V ) b◦ v(V ) b◦ v(V ) b v(V ) B~ (7.65) − h | i j 1 ··· i ··· j ··· k | i 1≤Xi

k dΩ(V , , V )= ( )i Σ T(v ) b(v ) b◦(v ) b(v ) B~ 1 ·· k − h | i 1 ··· i ··· k | i Xi=1 b b (7.67) + ( 1)i+j Σ b v , v b(v ) b◦(v ) b◦(v ) b(v ) B~ , − h | i j 1 ··· i ··· j ··· k | i 1≤Xi

Σ b(v ) b(v )(Q(1) + Q(n)) B~ , (7.68) h P | 1 ··· k ··· | i and move the sum of BRST operators all the way to the left using (7.22), until the sum of BRST operators hits the bra Σ and annihilates it. In doing so we obtain h | k ( )k−i Σ b(v ) b◦(v )T(v ) b(v ) B~ , (7.69) − h P | 1 ··· i i ··· k | i Xi=1 where the sign factor arises because Q was commuted through (k i) antighosts. We must − now move the T’s all the way to the left.P In doing so we obtain two type of terms; terms where

88 T’s have reached the bra Σ h |

k ( )k ( )i Σ T(v )b(v ) b◦(v ) b(v ) B~ , (7.70) − − h P | i 1 ··· i ··· k | i Xi=1 and terms arising from the commutators of T with b (see (7.21))

( )k ( )i+j Σ b([v , v ]) b(v ) b◦(v ) b◦(v ) b(v ) B~ , (7.71) − − h P | i j 1 ··· i ··· j ··· k | i 1≤Xi

7.4. Defining the String Field Multilinear Functions and Products

We have done all the preparatory work to be able to define precisely the string field multi- linear functions, and the string field products. We will first discuss the multilinear functions, give their definition and basis properties. These multilinear functions, when used for the dy- namical string field Ψ, give the string field vertices entering the string field actions. We show that these string field vertices are real. Then we define the string field products and prove some of their basic properties. We leave for 8 the proof of the main identity. § The String Multilinear Functions. The form Ω(0)g,n is of degree (6g 6+2n) and therefore B~ − can be integrated on the section σ( ) . Let = σ( ) denote the subspace of Mg,n ∈ Pg,n Vg,n Dg,n the section corresponding to the surfaces comprising string vertex. The string Dg,nb ⊂ Mg,n field vertex or multilinear function associated to this subset of surfaces is defined to be

(0)g,n B1, B2, , Bn = Ω . (7.72) ··· g Z B1···Bn  Vg,n

It is simply the integral over the subset of surfaces defining the string vertex, using the minimal area section. Thus given n string fields, the multilinear product gives us a number. This number, of course, is an ordinary number, complex, in general times a set of spacetime fields (recall the usual expansion of the string field in (3.1)).

⋆ Note that it would have been incorrect to naively identify the expressions in (7.70) and (7.71) with the first and second terms in equation (7.50).

89 As we discussed below Eqn.(7.37) the form entering Ω(0) entering the definition above B~ vanishes unless (G(B ) 2) = 0, and therefore, the multilinear function vanishes unless this i i − is the case (as statedP in (4.37)). Since the bra Σ entering Ω(0) is Grassmann even and the h | B~ number of antighost insertions is even, the statistics of the multilinear function is simply the statistics of the product of the B’s, in other words, the multilinear functions are intrinsically even (Eqn. (4.38)). It is also obvious from the above definition that the functions are indeed multilinear, in the form stated in Eqn. (4.35). It is convenient to be more explicit as far as the states to be inserted. Using (7.36) we are led to define Ω(0)g,n N Σ b(v ) b(v ), (7.73) h | ≡ g,nh P | 1 ··· 6g−6+2n which allows us to write (0)g,n g,n Ω = Ω B1 Bn . (7.74) B~ h | i···| i The bra Ωg,n is simply introduced for convenience of notation. It is a form on T ( ), but h | P Pg,n valued in the tensor product of n copies of the Hilbert space . It is not uniquely defined H b as such, since as we saw in the previous section, the various choices one has tob make affect Ωg,n , but do not affect the value of Ω(0)g,n due to the subsidiary conditions on the states h | B~ B . Whenever we will use Ωg,n it will appear contracted with suitable states. Using this | ii h | notation we write g,n B1, B2, , Bn = Ω B1 B2 Bn . (7.75) ··· g Z h | i| i···| i  Vg,n

Let us consider now the graded-commutativity property of the multilinear functions. By construction, the string vertices used to define the multilinear functions include a set of Vg,n Riemann surfaces with labeled punctures, with the condition (stated in 5.1 as condition (b)) § that the local coordinates are assigned in a labelling independent way, and if a surface is in R the set, the copies of with all other inequivalent labelings of the punctures are also included R in the set. This implies that the result of the multilinear product does not depend on which state is assigned to which puncture. Using the notation explained below Eqn. (7.36) this implies that

g,n g,n Ω (1, 2, , n) Bi Bj = Ω (1, 2, , n) Bi Bj , Z h ··· |···| i(i) ···| i(j) ··· Z h ··· |···| i(j) ···| i(i) ··· Vg,n Vg,n (7.76) and effectively means that the object

Ωg,n(1, 2, , n) , (7.77) Z h ··· | Vg,n

90 † acting on states, is symmetric under any exchange of labels of the Hilbert spaces. This means that the multilinear product simply picks up a sign corresponding to the statistics of the states that one commutes, that is

B , , B , B , , B =( )BiBi+1 B , , B , B , , B , (7.78) 1 ··· i i+1 ··· n g − { 1 ··· i+1 i ··· n}g  Note that the multilinear function we have introduced need always at least one argument, that is n 1, otherwise there is no state Σ to work with. In fact, since we did not introduce the ≥ h P | string vertices for n =0, 1, 2, so far we only have multilinear products defined for n 3 Vg=0,n ≥ when the genus is zero. When we studied the geometrical recursion relations we observed that the subsets for g 2 were all defined by the recursion relations. The case of g = 1 is Vg,0 ≥ peculiar in that the subset is not defined by the recursion relations. We will simply Vg=1,0 assume it is determined somehow (see 6.4). Let us therefore extend the definition of the § multilinear products to g 1 and n = 0. The idea is that the multilinear product, denoted in ≥ this case by must give us a number. Physically, this number should just be the result for {·}g the corresponding vacuum graph, but since the operator formalism does not give us a state for surfaces without punctures, we must use the same idea that was used in [36] to calculate partition functions. Consider then the subset , and note that since we have no punctures Vg,0 and therefore no local coordinates, this should be thought as a subset of . We then find Mg,0 a corresponding subset such that when we forget about the extra puncture we Dg,1 ∈ Mg,1 obtain the set . Then we define the form Ωg,0 on T ( ) by the expression Vg,0 e P~ Mg,0 Ωg,0(V~ , V~ )= Σ b(v ) b(v ) 1 , (7.79) 1 ··· 6g−6 h P | 1 ··· 6g−6 | i where the surface P σ( ) projects into the surface P~ as we forget about the ∈ Dg,1 ∈ Mg,0 puncture. Note that ghost number works out correctly, since, compared to the top form on e we have two less antighost insertions, but at the same time instead of inserting a state of Pg,1 ghost number two, we are inserting the vacuum, of ghost number zero (for g = 1 one must use b two antighost insertions). It was shown in [36] that this is a well defined form on . Now Mg,0 we simply set

= Ωg,0. (7.80) · g Z  Vg,0⊂Mg,0 For genus zero we will set =0. (7.81) · 0  The multilinear products for g = 0 and n =1, 2 are taken to be

B 0, and B , B B , QB . (7.82) 0 ≡ 1 2 0 ≡h 1 2i   The symmetry may not be manifest, but it must hold given the properties of the states representing the † surfaces. For example, the antighost insertions cannot be generically done treating the punctures in a symmetric way, however, it is clear that using Borel vectors the insertions can be moved around to show the symmetry of the bra Ωg,n . h |

91 Hermiticity of the String Field Vertices. Recall now from 4 that the string field vertices were § the key elements in constructing the string action. We had that

1 Sn(Ψ) Ψn . (7.83) g ≡ n! g  We want to show that given a dynamical string field Ψ satisfying the reality condition Ψ = | i h hc| Ψ imposed in 3.1, the string field vertex Ψn is real. The key geometrical input that −h | § { }g guarantees the hermiticity of the string field vertex is condition (c) in 5.1, which requires that § given a surface Σ then the mirror Σ∗ is also in , and that the local coordinates in Σ ∈ V V and Σ∗ be related by the antiholomorphic map that relates the two surfaces. The minimal area area problem that we use to define the local coordinates indeed gives us string vertices satisfying those conditions ( 6.4). In terms of the states representing the surfaces the above § conditions imply that Σ = Σ∗ , (7.84) h hc| h | which can be established by using the conserved charges method of the operator formalism [36]. In order to show that the string vertex is real we must consider the form we are integrating around the surface Σ (Σ) = N Σ b(v ) b(v ) Ψ~ , (7.85) g,nh | 1 ··· r | i and the form around the surface Σ∗

(Σ∗)= N Σ∗ b(v∗) b(v∗) Ψ~ , (7.86) g,nh | 1 ··· r | i v∗ where the Schiffer vector i is defined by the condition that the deformations it produces on ∗ the coordinates of Σ be the mirror images of the deformations the Schiffer vector vi produces on Σ. Under this condition, we must show that the number in (7.86) is just the complex conjugate of that in (7.85). This will imply the hermiticity of the string field vertex since we can then break up the total integral into two pieces one the hermitian conjugate of the other. The condition that the coordinates be related by the antiholomorphic map means that if the map takes a point p Σ in the domain of the local coordinate z into the point p∗ Σ∗, ∈ ∈ then the local coordinate z∗ satisfies

z∗(p∗)= z(p), (7.87)

where the bar will denote complex conjugation. Then given a Schiffer vector v(z) in Σ, de- forming z as z(p) z(p)+ v(z(p)), complex conjugating, we have → z(p) z(p)+ v(z(p)), → and using (7.87) it implies that

z∗(p∗) z∗(p∗)+ v(z∗(p∗)), (7.88) →

92 and we deduce that v∗(z)= v(z). More explicitly

vn ∗ vn v(z)= n−1 v(z)= v (z)= n−1 . (7.89) X z → X z This defines the Schiffer vectors v∗ entering in (7.86). Let us now begin by taking the hermitian conjugate of (7.85) to find

† (Σ) = N Σ b(v ) b(v ) Ψ~ , g,n h | 1 ··· r | i =( )nN Ψ~ [b(v )]† [b(v )]† Σ , − g,nh hc| r ··· 1 | hci (7.90) =( )dg,n N Ψ~ [b(v )]† [b(v )]† Σ , − g,nh | r ··· 1 | hci = N Ψ~ [b(v )]† [b(v )]† Σ , g,nh | 1 ··· r | hci where in the first step we got a factor of ( )n since there are n Hilbert spaces and each one − gives a minus sign due to (2.37). In the second step we used the reality condition on the string field, which gives us another ( )n, and the definition of the normalization factor in (7.37). − Finally, in the last step we rearranged the antighost insertions; since there are r = 2dg,n of them, this cancels the previous sign factor. The last equation can be written in the form

(Σ) = N Ψ~ M , (7.91) g,nh | i where the Grassmann even ket M is given by | i M = [b(v )]† [b(v )]† Σ . (7.92) | i 1 ··· r | hci It follows from (2.50) that (Σ) = N M Ψ~ , (7.93) g,nh | i and therefore we must simply evaluate

† † M = R ′ R ′ [b(v )] [b(v )] Σ . (7.94) h | h 11 |···h nn | 1 ··· r | hci Using the definition of b(v) in (7.19) and the expansion of the vector v in (7.89) we have

† b(v)= (bnv−n + bnv−n), and b(v) = (b−nv−n + b−nv−n), (7.95) X X and therefore using (2.58) we have

† Rii′ [b(i′)(v)] = Rii′ (b−n(i′)v−n + b−n(i′)v−n), h | h | X = R ′ (b v + b v ), (7.96) h ii | n(i) −n n(i) −n X∗ = R ′ [b(v )] . h ii | (i) Using this result, together with the Grassmann even property of Σ one finds that Eqn. | hci

93 (7.94) gives M = Σ b(v∗) b(v∗). (7.97) h | h hc| 1 ··· r Finally, using Eqn.(7.84) and substituting back into (7.93) we get

(Σ) = N Σ∗ b(v∗) b(v∗) Ψ~ = (Σ∗), (7.98) g,nh | 1 ··· r | i which is the result we wanted to establish. This concludes our proof that the vertices of the string field theory are hermitian. Having shown in 3.2 the hermiticity of the kinetic term, § this establishes the hermiticity of the full string field theory. String Field Products. Having defined the multilinear string field functions let us now define the closely related string products. These products take a set of string fields, and give us another string field. In the same way as the multilinear functions they are labeled by the genus g. We define

′ Φs g,n B1, , Bn−1 ( ) Ω Φs 0 Φs e B1 Bn−1 , (7.99) ··· g ≡ − · Z h | i | i | i···| i   Xs Vg,n e

where the sum extends over a basis of states Φs complete in the subspace of states annihilated by (L L ). The label e, standing for external, on the state Φ is the label for the Hilbert 0 − 0 | sie space of the resulting product. This state is not contracted with Ωg,n . The Hilbert spaces of e h | the bra Ωg,n are labeled by 0, 1, , n 1. The state Φ is contracted with the 0-th Hilbert h | ··· − | si0 space, and each state B is contracted with the i-th Hilbert space. One easily rewrites the | ii product in the following forms

′ Φs g,n B1, , Bn−1 = ( ) Φs e Ω Φs 0 B1 Bn−1 , ··· g − | i · Z h | i | i···| i   Xs e Vg,n ′ = ( )Φs Φ Φ , B , , B , − | sie · s 1 ··· n−1 g Xs  ′ e (7.100) = Ωg,n b−(e) Φ Φc B B , h | 0 | si0| sie| 1i···| n−1i Xs Z Vg,n

g,n ′ = Ω R 0e B1 Bn−1 . Z h | i| i···| i Vg,n f These are obtained from (7.99) as follows. In the first relation we simply used the facts that Ωg,n is Grassmann even and that Φ and Φ commute with each other. In the second form h | | si | si we made use of the definition of the multilinear functions. In the third form we made explicit e the b− factor in the definition of Φ . In the last form we simply recognized the form of the 0 | si reflector. Multilinearity of the product is manifest from the definition. Moreover it is also e clear that the product is intrinsically Grassmann odd, because the multilinear functions are

94 Grassmann even and the sewing ket R′ entering the definition of the product is Grassmann | i odd. Finally the multilinear products are clearly graded commutative just as the multilinear f functions. The multilinear functions and the multilinear products are simply related to each other via the linear inner product. In fact, using the last expression in (7.100) we have

−(e) g,n −(e) ′ A, B1, , Bn−1 = e A c Ω b R 0e B1 Bn−1 , ··· g h | 0 · Z h | 0 | i| i···| i  Vg,n

g,n −(e) −(e) ′ = Ω e A c b R 0e B1 Bn−1 , Z h | h | 0 0 | i | i···| i Vg,n 

g,n (7.101) = Ω A R0e B1 Bn−1 , Z h |h | i| i···| i Vg,n

g,n = Ω A B1 Bn−1 , Z h | i| i···| i Vg,n = A, B , , B . 1 ··· n g 

Let us conclude this section by establishing two properties (given in (4.28) and (4.29) )that arise because of symmetry properties of the states. The first one is

′ , Φs Φs, =0, (7.102) ··· ··· g1 g2 Xs     e where the dots denote arbitrary string fields. This expression amounts to

′ g1,k ′ ′ g2,m ′ ′ Ω (0, k, l ) R Φ Ω (l,k ) R ′ Φ ′ (7.103) h ··· | 0ei···| si(k) · h ··· | ll i| si(k ) ··· Xs Z Z Vg ,n Vg ,m 1 f 2 f e

Φs pushing the state Φ in front of the state Φ ′ we get a sign factor ( ) (since the | si(k) | si(k ) − operator in between is odd) and we recognize the sewing ket e

′ g1,k ′ ′ g2,m ′ ′ ′ Ω (0, k, l ) R Ω (l,k ) R ′ R ′ . (7.104) h ··· | 0ei··· h ··· | ll i| kk i ··· Xs Z Z Vg ,n Vg ,m 1 f 2 f f

′ ′ This term vanishes because the product of kets R ′ R ′ is antisymmetric under the Hilbert | ll i| kk i spaces label exchange k l′; k′ l, but the rest of the expression in (7.104) is symmetric ↔ ↔ f f

95 under this exchange. For the case of the identity

′ ( )Φr ( )Φs Φ , Φ , , Φ , Φ , =0. (7.105) − − s s ··· r r ··· g Xr,s   e e the left hand side amounts to

g,n ′ ′ ′ ′ Ω ( ,l,l , ,k,k ) R ll′ R kk′ , (7.106) Z h ··· ··· ··· |···| i| i··· Vg,n f f and this clearly vanishes for the same reason as in the case above.

8. Establishing the Main Identity

In sections 5 and 6 we learned how to find subsets , called string vertices, satisfying § § Vg,n a nontrivial set of recursion relations and conditions. The reason they satisfy such relations is that sewing of the string vertices with the standard Feynman combinatorial rules gives a single cover of all relevant moduli spaces. In 7 we learned how to define string products as § integrals of forms over the subsets. In the present section we will use the recursion relations Vg,n satisfied by the string vertices in order to prove the identity satisfied by the string products. This identity, called the main identity, and given in (4.10), guarantees the consistency of the string field theory, as we have established in 4. It relates the failure of the BRST operator to § act as a derivation on the string products to the failure of the string products to satisfy Jacobi type identities. We begin our work with the expression for the violation of the derivation property

n (I) Q B , , B + ( )(B1+···Bi−1) B , QB , , B , (8.1) ≡ 1 ··· n g − 1 ··· i ··· n g   Xi=1   and now using the expression for the string field product given in the last right hand side of Eqn. (7.100) we write (I) as

n g,n+1 ′ (B1+···Bi−1) ′ (I) = Ω Qe R 0e B1 Bn + ( ) R 0e B1 Q Bi Bn , Z h | | i| i···| i − | i| i··· | i···| i Xi=1 Vg,n+1 f f n g,n+1 ′ = Ω Q0 + Qi) R 0e B1 Bn , − Z h | | i| i···| i Xi=1 Vg,n+1 f n ′ Φs+1 g,n+1 = ( ) Ω Qi Φs 0 Φs e B1 Bn , − Z h | | i | i | i···| i Xs Xi=0  Vg,n+1 e (8.2) where use was made of the fact that Ω is Grassmann even to commute Q through it, and in h | e the second step we used the BRST property of the sewing ket (Eqn.(2.77)) and brought all the

96 BRST operators to the same position. In order to be able to use the BRST properties of the forms derived earlier it is convenient to break up the sewing ket and take out of the integrand the state with external label. Due to the presence of the BRST operators we get an extra sign factor of ( )Φs+1 and therefore we obtain − ′ (I) = Φ Ωg,n+1 Q ) Φ B B , | sie · h | i | si0| 1i···| ni Xs Z X Vg,n e +1 (8.3) ′ (0)g,n+1 = Φs e Ω , | i · Z ( Q)ΦsB1···Bn Xs e Vg,n+1 P

where we used our standard notation for forms in . We can now use the BRST property Pg,n+1 (7.49) and Stokes theorem to obtain: b

′ (−1)g,n+1 ′ (−1)g,n+1 (I) = Φs e dΩ = Φs e Ω . (8.4) | i · ΦsB1···Bn | i · ΦsB1···Bn Xs Z Xs Z e Vg,n+1 e ∂Vg,n+1 But now recall that in 5 it was seen that (Eqn. (5.2)) ∂ = ∂ R , where R denotes § Vg,n+1 − p 1 1 the Feynman graphs with one propagator, and ∂p, the propagator boundary, that is the set of surfaces obtained by gluing two elementary vertices via the sewing condition zw = t with t ∂D, or by gluing two punctures in the same elementary vertex with the same sewing ∈ condition. These two alternatives are illustrated in Eqn. (5.1). Since we must keep track of the orientation of the manifolds where we integrate, the sign factor relating the vertex boundary to the Feynman diagram boundary implies that (8.4) becomes

′ (−1)g,n+1 (I) = Φs e Ω . (8.5) − | i · ΦsB1···Bn Xs Z e ∂pR1

We have to understand now the behavior of the form Ω(−1) appearing in the above equation on the set of surfaces ∂pR1. To this end we break up the set of surfaces into the its two components tree loop ∂pR1 = R1 + R1 , (8.6) tree loop where R1 denote the surfaces obtained by gluing two string vertices and R1 denote the surfaces obtained by gluing two punctures on a single string vertex. It follows from (8.6) that Eqn. (8.5) can be written in the form

′ (−1)g,n+1 (−1)g,n+1 (I) = Φs e Ω + Ω . (8.7) − | i ·  ΦsB1···Bn ΦsB1···Bn  Xs Z Z Rtree loop e 1 R1

We now claim that the following ‘factorization’ equations hold

97 ′ (−1)g,n+1 Φr+Φs (0)g1,l+2 (0)g2,k+1 Ω = ( ) σ(il, jk) Ω Ω , (8.8) ΦsB1···Bn ΦsBi ···Bi Φr ΦrBj1 ···Bjk Z − g +g =g Z 1 l · Z tree Xr 1X2 R1 {il,jk};l,k≥0 Vg1,l+2 Vg2,k+1 l+k=n e for the surfaces of the ‘tree’ type, and for the surfaces of the ‘loop’ type we get

1 ′ Ω(−1)g,n+1 = ( )Φr+Φs Ω(0)g−1,n+3 . (8.9) ΦsB1···Bn Z 2 − Z ΦsΦrΦrB1···Bn loop Xr R Vg−1,n+3 1 e Let us assume for the time being the correctness of the factorization equations above and complete the proof of the main identity. It follows by linearity that Eqn. (8.8) can be written as

(−1)g,n+1 Φs (0)g1,l+2 Ω = ( ) σ(il, jk) Ω , (8.10) ΦsB1···Bn ΦsBi1 ···Bil X Z g +g =g − Z tree 1X2 R1 {il,jk};l,k≥0 Vg1,l+2 l+k=n where the state X is given by | i ′ X = ( )Φr Φ Ω(0)g2,k+1 = B B , (8.11) r ΦrBj ···Bj j1 jk | i − | i· 1 k ··· g2 r Z X V   e g2,k+1 where use was made of Eqn.(7.100). Using now the definition of the multilinear product in (7.72) and incorporating the value of the ket X , Eqn. (8.10) gives us | i

(−1)g,n+1 Φs Ω = ( ) σ(i , j ) Φs, Bi Bil , Bj Bjk , (8.12) ΦsB1···Bn l k 1 1 g2 g1 Z g +g =g − ··· ··· tree 1X2    R1 {il,jk};l,k≥0 l+k=n

In a similar way the ‘loop’ factorization equation (8.9) can be written as

′ (−1)g,n+1 1 Φr+Φs Ω = ( ) Φs, Φr, Φr, B1, , Bn . (8.13) Z ΦsB1···Bn 2 − ··· g−1 loop Xr  R1 e

Substitute now the results of the last two equations into (8.7)

′ Φs+1 (I) = ( ) Φs e σ(il, jk) Φs, Bi1 Bil , Bj1 Bjk − | i ·  ··· ··· g2 g1 Xs g1X+g2=g    e {il,jk};l,k≥0 l+k=n (8.14) 1 ′ + ( )Φr Φ , Φ , Φ , B , , B , 2 − s r r 1 ··· n g−1 Xr  e

98 but using the definition of the string product (Eqn.(7.100)) we find ′ 1 Φr (I) = σ(il, jk) Bi1 , , Bil , Bj1 , , Bjk ( ) Φr, Φr, B1, , Bn . − ··· ··· g2 g1 − 2 − ··· g−1 g1X+g2=g     Xr   {il,jk};l,k≥0 e l+k=n (8.15) Comparing with Eqn. (8.1) we have succeeded in showing that n 0 = Q B , , B + ( )(B1+···Bi−1) B , QB , , B 1 ··· n g − 1 ··· i ··· n g   Xi=1  

+ σ(il, jk) Bi1 , , Bil , Bj1 , , Bjk ··· ··· g2 g1 (8.16) g1X+g2=g     {il,jk};l,k≥0 l+k=n 1 ′ + ( )Φr Φ , Φ , B , , B , 2 − r r 1 ··· n g−1 Xr   e which is the desired result. In order to complete the proof we must establish the factorization relations that were assumed in the above derivation.

Proving the Factorization Equations. Let us begin with the ‘tree’ type factorization equation. Consider the form Ω(−1) for this case (−1)g,n+1 Σ Σ Σ Σ Ω (V , V )= Ng,n+1 Σ b(v ) b(v ) Φs B1 Bn , (8.17) ΦsB1···Bn 1 ··· 6g−6+2n+1 h | 1 ··· 6g−6+2n+1 | i| i···| i where the stateb Σ bcorresponds to a Riemann surface Σ Rtree. In order to emphasize the h | ∈ 1 fact that the Schiffer variations are done on the punctures of the surface Σ we have included the superscript on the tangent vectors and on the associated Schiffer variations. A surface Σ Rtree is built by sewing, so it actually determines two surfaces Σ and Σ ∈ 1 1 2 with coordinate curves around their punctures. This means that there is a phase ambiguity b b for the local coordinates, in particular for the local coordinates around the punctures that are to be sewn. This ambiguity makes the sewing parameter of the glued surface ambiguous. If we make a choice of marked points around the punctures of the surfaces Σ1 and Σ2 we obtain the surfaces Σ and Σ respectively, and we can determine the sewing parameter t. We write 1 2 b b Σ Rtree Σ = Σ Σ , with Σ and Σ , (8.18) ∈ 1 → 1 ∪t 2 1 ∈ Vg1,l+2 2 ∈ Vg2,k+1 with l,k 0, l +k = n, and where denotes sewingb with sewing parameterb t = exp(iθ). Note ≥ ∪t that the surface Σ1 has been assumed to contain the puncture corresponding to the state Φs,

the l punctures corresponding to l states Bil , and one extra puncture r which is sewn to the ′ | i puncture r of the surface Σ2. This latter surface contains k additional punctures. The sewing statement implies that the states representing the surfaces Σ, Σ1, and Σ2 satisfy the relation

(r) (r) L (r) (r) L0 0 iθ(L0 −L0 ) θ Σ = Σ Σ t t R ′ = Σ Σ e R ′ = Σ Σ R ′ . (8.19) h | h 1|h 2| | rr i h 1|h 2| | rr i h 1|h 2| rr i Recall that the operator formalism construction of a state associated to a surface requires the complete specification of coordinates around the punctures. Since we only have (naturally)

99 coordinates defined up to a phase, to construct Σ , for example, we specify arbitrarily the h | marked points, this choice being irrelevant since the states that are to be contracted with Σ h | satisfy the L L = 0 condition. What we have noted is that the phase ambiguity around the 0 − 0 local coordinates to be sewn must be fixed, in order to be able to define the sewing parameter. The phases around the other punctures of the constituent surfaces Σ1 and Σ2 are irrelevant. The surface Σ with unmarked coordinate curves will be denoted as Σ. tree We have to consider integration over the set of surfaces in Rb1 . This space can be considered as a base space with a U(1) fiber, representing the sewing angle. Vg1,l+2 × Vg2,k+1 We will split the parameters of integration into a group of parameters defining the surface Σ1, a group of parameters defining the surface Σ2 and the sewing parameter. Then we will see that the integrand factorizes, and we shall be able to complete the proof of factorization. The U(1) fiber is generated by deformations of the surface Σ associated with the sewing parameter. From Σ we determine uniquely two surfaces Σ1 and Σ2 with coordinate curves. We then make a choice of marked points around the punctures to be sewn and obtain the surfaces b b Σ , Σ and the sewing parameter t. We let V Σ denote the tangent (in T ( )) generating the 1 2 θ Σ P deformation ∂ , and let vΣ denote the corresponding Schiffer variation. This Schiffer variation ∂θ θ b b b is supported on the neighborhood of the punctures of Σ, thus in particular, away from the region of sewing. It is clear from the fact that any tangent can be represented by a Schiffer variation vΣ that θ exists, and is defined uniquely up to a Borel vector. This deformation, however, can be produced directly by changing the coordinates of the sewing ket by the corresponding phase rotation. We must therefore have that

Σ (r) θ Σ T(v )= Σ Σ i(L L ) R ′ . (8.20) h | θ h 1|h 2| 0 − 0 | rr i Since the connection conditions for the stress energy tensor are the same as those for the antighost field, the above equation implies that we must have

Σ (r) θ Σ b(v )= Σ Σ i(b b ) R ′ . (8.21) h | θ h 1|h 2| 0 − 0 | rr i

Let us now show how to obtain from a deformation of the surface Σ1 a deformation of the surface Σ. Again, from Σ we determine uniquely two surfaces Σ and Σ , and then (by 1 b 2 a choice of marked points) the surfaces Σ , Σ and the sewing parameter t. Now consider a 1 2 b b Σ1 deformation of the surface Σ1, that is, a tangent vector V T ( ). We choose a tangent i ∈ Σ1 P b vector V Σ1 T ( ), whichb projects down to V Σ1 as web forget aboutb the phases of the local i ∈ Σ1 P i b coordinates. We use this tangent vector V Σ1 tob deform the surface Σ , and then we sew i b 1 the deformed Σ1 to Σ2 with sewing parameter t to get a deformed Σ surface. Let us denote Σ the tangent that generates this deformation of Σ by V (i, Σ1), where the information inside the parenthesis indicates that this tangent originated from the i-th deformation of Σ1. The Σ b vector V (i, Σ1) is not uniquely determined, due to the choices made at several places. The ambiguity, however, is simple. Different choices would result in the addition to V Σ(i, Σ ) ofa b 1 b 100 Σ vector along Vθ , which, as mentioned in the previous paragraph, corresponds to a change in the sewing parameter. Denote the associated Schiffer variations by b V Σ1 vΣ1 and V Σ(i, Σ ) vΣ(i, Σ ). (8.22) i ↔ i 1 ↔ 1 b vΣ1 While the vector i is supported on all the punctures of Σ1, including the one to be sewn, Σ the vector v (i, Σ1) is supported around the punctures of Σ. The above discussion relating the deformations implies that

Σ Σ1 θ Σ T(v (i, Σ )) = Σ Σ T(v ) R ′ . (8.23) h | 1 h 1|h 2| i | rr i Again, a similar equation must hold for antighosts:

Σ Σ1 θ Σ b(v (i, Σ )) = Σ Σ b(v ) R ′ . (8.24) h | 1 h 1|h 2| i | rr i

The index i runs from one up to d1, where d1 =2dg1,l+2 is the (real) dimensionality of the set . A completely analogous discussion holds for the component Σ of the surface. In this Vg1,l+2 2 case the index i runs from one up to d2, where d2 = 2dg2,k+1 is the (real) dimensionality of . Vg2,k+1 We can now consider the form Ω(−1) of Eqn. (8.17), with input tangent vectors the vectors defined by the deformations of Σ1, Σ2 and the sewing angle

(−1)g,n+1 Σ Σb b Σ Σ Σ Ω V (1, Σ1), V (d1, Σ1) ; V (1, Σ2), V (d2, Σ2) ; V ΦsB1···Bn ··· ··· θ Σ Σ Σ  Σ = Ng,n+1 Σb1 Σ2 b(v (1b, Σ1)) b(vb (d1, Σ1)) bb(v (1, Σ2)) b b(v (d2, Σ2)) (8.25) h |h | ··· ··· Σ θ    b(v ) R ′ Φ B B . · θ | rr i| si| 1i···| ni Note that the number of antighost insertions associated to each of the surfaces is even. We now can use Eqns.(8.21) and (8.24) to rewrite the right hand side as

Φs b vΣ1 b vΣ1 =( ) Ng,n+1 Σ1 ( 1 ) ( d ) Φs − h | ··· 1 | i (8.26) Σ2 Σ2 −(r) θ Σ2 b(v ) b(v )(ib R ′ ) B1 Bn , ·h | 1 ··· d2 0 | rr i | i···| i where the sign factor appeared by commuting the state Φ across an odd number of antighosts. | si We now rearrange the states to find

Φs b vΣ1 b vΣ1 =( ) σ(il; jk)(iNg,n+1) Σ1 ( 1 ) ( d ) Φs Bi1 Bil − h | ··· 1 | i| i···| i (8.27) Σ2 Σ2 −(r) θ Σ2 b(v ) b(v )(b R ′ ) Bj Bj , · h | 1 ··· d2 0 | rr i | 1 i···| k i where the sign factor σ is the sign involved in rearranging the sequence (b , B , B ) into 0 1 ··· n the sequence (B B ; b ; B , B ). Let us now expand the sum over states involved in i1 ··· il 0 j1 ··· jk

101 the reflector to obtain

Φr +Φs b vΣ1 b vΣ1 = ( ) σ(il; jk)(2πiNg,n+1) Σ1 ( ) ( ) Φs Bi1 Bil Φr − h | 1 ··· d1 | i| i···| i| i Xr e (8.28) Σ2 Σ2 1 − Σ2 b(v ) b(v ) exp(iθL ) Φr Bj Bj , ·h | 1 ··· d2 2π 0 | i | 1 i···| k i   − where in the sum one includes states that are not annihilated by L0 . We now use the relation

(2πi)Ng,n+1 = Ng1,l+2 Ng2,k+1, (8.29) which follows from the definition of N in (7.37) and

dg,n+1 =1+ dg1,l+2 + dg2,k+1, (8.30) which is the familiar statement that the ingredient spaces on the left and on the right of the geometrical equation differ by complex dimension one. Equations (8.28) and (8.29) imply that

(−1)g,n+1 Σ Σ Σ Σ Σ Ω V (1, Σ1), V (d1, Σ1) ; V (1, Σ2), V (d2, Σ2) ; V ΦsB1···Bn ··· ··· θ  b b b b b Φr +Φs (0),g1,l+2 Σ1 Σ1 (0),g2,k+1 Σ2 Σ2 = ( ) σ(il; jk) Ω (V , V ) Ω 1 − (V , V ). (8.31) 1 d1 [ exp(iθL )Φr]Bj ···Bj 1 d2 − ΦsBi1 ···Bil Φr ·· · 2π 0 1 k ·· Xr b b b b b b b b Note that we have obtained a relatione between forms defined on different spaces. The best way of thinking about the above equation is that given the sets of tangents in T and T we have Σ1 Σ2 shown that the right hand side equals the form on the left hand side evaluated on associated b b tangents in TΣ. It is necessary that whatever choices we made to get the tangents in TΣ the result of the left hand side is the same. This is clear, because as we discussed before, the b Σ b ambiguity amounts to changing the tangents in TΣ by vectors along Vθ . Thus, for different choices Eqn.(8.23) would read b b Σ Σ Σ1 θ Σ T(v (i, Σ )) + T(v ) = Σ Σ T(v ) R ′ . (8.32) h | 1 θ h 1|h 2| i | rr i   The related equation for the antighosts would read

Σ Σ Σ1 θ Σ b(v (i, Σ )) + b(v ) = Σ Σ b(v ) R ′ , (8.33) h | 1 θ h 1|h 2| i | rr i   and if we used this in relating Eqns.(8.26) and (8.25) we would find that the extra terms vanish b vΣ tree due to the presence of ( θ ) in (8.25). We are now essentially done. In integrating over R1 we must consider all possible splittings, as indicated in the geometrical equation (5.1). Thus we have 1 = , (8.34) Z 2 g +g =g Z · Z · Z tree 1X2 1 R1 {il,jk};l,k≥0 Vg1,l+2 Vg2,k+1 S l+k=n where S1 is the integral over the sewing angle. Since we always want to consider the state Φs as insertedR on the surface which is said to have genus g1, the terms in the geometrical equation

102 when Φs ends up in the surface of genus g2 are relabeled and simply cancel the factor of 1/2 in the above equation. Integrating equation (8.31) using (8.34) we have

(−1)g,n+1 Φr +Φs (0)g1,l+2 (0)g2,k+1 Ω = ( ) σ(i , j ) Ω Ω dθ − , ΦsB1···Bn l k ΦsBi ···Bi Φr [ 2π exp(iθL0 )Φr]Bj1 ···Bjk Z − g +g =g Z 1 l · Z tree Xr 1X2 R1 {il,jk};l,k≥0 Vg1,l+2 Vg2,k+1 R l+k=n e (8.35) where the integral over the sewing parameter went into the state in the second form. This − integral simply projects to the subspace of states annihilated by L0 . This reduces the sum ′ over all states r to the usual sum r, whereupon the above equation is precisely the first factorization equation. P

We must now deal with the case of the ‘loop’ factorization equation. Given a surface Σ Rloop of genus g and n punctures, by cutting the surface along the sewing line, and ∈ 1 restoring back the semiinfinite cylinders, we obtain the surface Σo, where the ‘o’ stands for open (the surface was opened up) and the hat indicates that there is no canonical definition b of the phase of the local coordinates around the two punctures in question. We then make a choice of the phases by marking arbitrary points in the coordinate curves associated to the two punctures, and obtain the surface Σo (no hat). At this stage the sewing parameter t corresponding to the surface Σ can be determined. We therefore have

(r) (r) L (r) (r) L0 0 iθ(L0 −L0 ) θ Σ = Σ t t R ′ = Σ e R ′ = Σ R ′ . (8.36) h | h o| | rr i h o| | rr i h o| rr i

Again, our concern is the form Ω(−1) that reads

(−1)g,n+1 Σ Σ Σ Σ Ω (V , V )= Ng,n+1 Σ b(v ) b(v ) Φs B1 Bn , (8.37) ΦsB1···Bn 1 ··· 6g−6+2n+1 h | 1 ··· 6g−6+2n+1 | i| i···| i b b where now Σ Rloop. This time the 6g 6+2n + 1 dimensional region of integration, ∈ 1 − corresponding to that number of antighost insertions in the above, must be related to an integration over , which is of dimension 6g 6+2n, plus an integration over the Vg−1,n+3 − sewing angle. Just as we did in the ‘tree’ case, we now have that

Σ (r) θ Σ T(v )= Σo i(L0 L0) R ′ , h | θ h | − | rr i (8.38) Σ (r) θ Σ b(v )= Σ i(b b ) R ′ , h | θ h o| 0 − 0 | rr i

vΣ Σ where θ is the Schiffer variation corresponding to the tangent vector Vθ that generates the deformation ∂/∂θ. Also (in a way analogous to the way we discussed the other deformations b Σo in the tree case) for a given Σ once we fix Σo and the sewing parameter, a tangent Vi can be used to choose a tangent V Σo , and this tangent is used to deform the surface Σ , whichb is then i o b

103 Σ sewn back to give a deformed Σ, thus defining a tangent V (i, Σo). This tangent is ambiguous up to a vector along the tangent V Σ. We therefore have, denoting the Schiffer variations by θ b b V Σo vΣo and V Σ(i, Σ ) vΣ(i, Σ ), (8.39) i ↔ i o ↔ o b that Σ Σo θ Σ T(v (i, Σo)) = Σo T(v ) R ′ , h | h | i | rr i (8.40) Σ Σo θ Σ b(v (i, Σ )) = Σ b(v ) R ′ . h | o h o| i | rr i Letting d 2d =6g 6+2n, we then have that Eqn.(8.37) can be evaluated for the o ≡ g−1,n+3 − tangents discussed above giving

(−1)g,n+1 Σ Σ Σ Ω V (1, Σo), V (do, Σo) ; V ΦsB1···Bn ··· θ (8.41) b vΣ b vΣ  b vΣ θ = Ng,n+1 Σbo ( (1, Σob)) ( (dbo, Σo)) ( ) R ′ Φs B1 Bn . h | ··· θ | rr i| i| i···| i   We now can use Eqns. (8.38) and (8.40) to rewrite the right hand side as

Φs Σo Σo −(r) 1 − c = ( ) (2πiNg,n+1) Σo b(v ) b(v ) Φs b exp(iθL ) Φr Φ B1 Bn , − h | 1 ··· do | i 0 2π 0 | i | ri| i···| i Xr   (8.42) where the sign factor appeared by moving the state Φ through b−. Using the form notation | si 0 (and the same observations as we made before with the normalization factor) we find

(−1)g,n+1 Σ Σ Σ Ω V (1, Σo), V (do, Σo) ; V ΦsB1···Bn ··· θ  b b b = ( )Φr+Φs Ω(0)g−1,n+3 V Σo , V Σo . (8.43) 1 − 1 do − Φs[ 2π exp(iθL0 )Φr]ΦrB1···Bn ··· Xr  e b b loop The integral over R1 can be written as

1 = , (8.44) Z 2 Z · Z loop 1 Vg−1,n+3 S R1 and using this on (8.43) we find

1 Ω(−1)g,n+1 = ( )Φr+Φs Ω(0)g−1,n+3 , (8.45) ΦsB1···Bn dθ − Z 2 − Z Φs[ 2π exp(iθL0 )Φr]ΦrB1···Bn loop Xr R Vg−1,n+3 R 1 e and upon doing the integral over θ the sum over states reduces to the sum over states satisfying − the L0 constraint, and the above equation becomes precisely the ‘loop’ factorization equation we wanted to establish. All in all this concludes our proof of the main identity.

104 9. Algebraic Structures on the BRST Cohomology

In this section we wish to connect our study of the algebraic structure of both the classical and quantum closed string field theory to the algebraic structures derived in the context of two-dimensional closed string theory. It was observed by Witten and the author [17] that the tree-level Ward identities of the theory could be neatly summarized as a homotopy Lie algebra on the BRST cohomology, with products m , m , m ... E. Verlinde [18] showed that the 2 3 4 ··· complete higher genus Ward identities could be obtained from a generating function satisfying a Batalin-Vilkovisky type master equation (for applications of the Ward identities see also Ref. [53]). In both works the possible relation to similar structures in closed string field theory was pointed out. In this section we derive the connection, by showing how to obtain from the basic identity relating the products in string field theory the structures on cohomology. We have shown in 6.4 that the string vertices have ‘stubs’ of length π around each punc- § ture. These stubs prevent the appearance of curves that are shorter than 2π when vertices are sewn together. Accordingly the string vertices include all surfaces whose internal Vg,n foliations are all of height smaller than 2π. We observed that the length of the stubs can be changed at will, as long as they are kept longer than π. For stubs of length l the string vertices (l) include all surfaces whose foliations are all of height smaller than 2l. The important Vg,n thing to realize is that as we let the parameter l grow every set becomes larger and larger. Vg,n This happens because (l) (l′) for l < l′, which in turn, is manifest from the above Vg,n ⊂ Vg,n description of the ’s. Thus, the sets (l), as l increases, are becoming smaller V Mg,n − Vg,n and smaller neighborhoods of the compactification divisor of (the set of noded surfaces Mg,n in ). As l these sets go simply into the noded surfaces. Equivalently, as l Mg,n → ∞ → ∞ (the moduli space without compactification, or nodes). One may think, at first Vg,n →Mg,n glance, that the growth of all vertices is incompatible with the fact that the Feynman graphs produce smaller and smaller neighborhoods of the compactification divisor, but this is clearly wrong, as the above discussion makes manifest. Indeed, think of the limit l , when the → ∞ propagator simply joins surfaces through a node. The nodes of actually correspond to Mg,n the sets with g + g = g; n + n = n + 2, plus the set ; namely, Mg1,n1 × Mg2,n2 1 2 1 2 Mg−1,n+2 the nodes include full copies of the lower dimensional compactified moduli spaces. Therefore, given that the propagator just joins surfaces through nodes, the lower vertices must all Vg,n coincide with the spaces in order for the Feynman graphs to produce the whole set of Mg,n noded surfaces. Consider the structure of the multilinear functions , constructed explicitly in 7. {···}g § They give, for a given set of of n states B , the integral over , with the Polyakov measure, i Vg,n of the correlation function of those states. Thus, they simply give us the part of the string amplitude corresponding to the subset of . Therefore as l , and as a consequence Vg,n Mg,n →∞ (l) , the multilinear product goes into the full string amplitude for the states B , Vg,n →Mg,n i associated to the corresponding genus (Of course, this assumes that for the given choice of external momenta, or parameters, the amplitude is finite, since otherwise the infinite part would indeed come from the neighborhoods of the noded surfaces, and the limits would not make sense).

105 Consider the following sum of multilinear functions in which each term has a BRST oper- ator acting on a state

QB , B , , B +( )B1 B , QB , , B + +( )B1+···Bn−1 B , , QB , (9.1) 1 2 ··· n g − 1 2 ··· n g ··· − 1 ··· n g    and the multilinear functions now correspond to integrals over the full moduli space . Mg,n This can be rewritten as

n i−1 Bk (0)g,n (−1)g,n ( ) k=0 B1, , QBi, , Bn = Ω = dΩ , − P ··· ··· g Z ( Q)B1···Bn Z B1···Bn Xi=1  Mg,n P Mg,n (9.2) = Ω(−1)g,n, Z B1···Bn ∂Mg,n where we used the definition of the multilinear functions, the BRST property of the forms, and Stokes theorem. We now argue that the final expression should be taken to be zero

n i−1 Bk ( ) k=0 B1, , QBi, , Bn =0. (9.3) − P ··· ··· g Xi=1 

Basically, what we are saying is that if the original integrand is well defined in (rather Mg,n than just ), that is, the integrand does not diverge at the degenerate surfaces, then the Mg,n original integral could have been done over , and since this space has no boundary we Mg,n do not get boundary terms. This is the essence of BRST decoupling, it is clearly a formal statement since one must check nice behavior near degeneration (Making sure this holds can require analytic continuation on external parameters, and typically at the quantum level, correcting the two-dimensional conformal field theory). We will assume Eqn. (9.3) holds. The usual statement of BRST decoupling is that

Φ , B , B , B =0, (9.4) t 2 3 ··· n g  where Φ denotes a BRST trivial state, namely, Φ = Q α , and all the B states, with t | ti | i | ii i 2, are annihilated by the BRST operator. This equation follows from (9.3) if we take ≥ B = α. Moreover, for arbitrary states α and β , and states B annihilated by Q for i 3, 1 | i | i | ii ≥ we also have Qα, β, B , , B +( )α α, Qβ, B , B =0, (9.5) 3 ··· n g − 3 ··· n g   which follows from (9.3) by taking B1 = α and B2 = β. The two equations above will be of utility to us later.

We now consider choosing a suitable basis of states. The states Φs of the conformal − − | i field theory, satisfying the b0 and L0 constraints will be decomposed into physical states Φp, or semi-relative cohomology classes (states annihilated by Q, that are not Q on something),

106 trivial states Φ (states that are of the type Q on something) and unphysical states Φ | ti | ui (states that are not annihilated by Q). The physical states are ambiguous up to trivial states, and the unphysical states are ambiguous up to the addition of physical and trivial states. We choose a basis of states such that the unphysical states, are all orthogonal to each other and orthogonal to the physical states. Moreover, since the trivial states must be all orthogonal to each other and to the physical states; upon choosing this basis we have that the states conjugate to physical states must be physical states and that the trivial and unphysical states are conjugates to each other. A detailed discussion on how to arrive to such basis has been given in Appendix C of the second paper in Ref.[14]. Given an unphysical state Φ denote | ui i the associated conjugated state by Φ ′ , that is | t i i

c Φ = Φ ′ , (9.6) | ui i | t i i we then define

− c − Φ = b Φ = b Φ ′ Φ . (9.7) | ui i 0 | ui i 0 | t i i≡| ti i e Therefore we have that

Φ = Φ , and Φ = Φ , (9.8) | ui i | ti i | ti i −| ui i e ⋆ e where the second relation is readily established. Using this basis of states, an arbitrary string product can be written in the form (using the second equation in (7.100))

′ B , B , , B = ( )Φs Φ Φ , B , , B 1 2 ··· n g − | si· s 1 ··· n g   Xs  e = ( )Φp Φ Φ , B , , B − | pi· p 1 ··· n g Xp  e Φu + ( ) i Φ Φ , B , , B + Φ Φ , B , , B . − | ti i· ui 1 ··· n g | ui i· ti 1 ··· n g Xi   (9.9) where two minus signs (the first from the opposite statistics of Φui and Φti , and the second from (9.8)) cancelled in the last term of the right hand side. We can now begin our reduction of the main identity (4.10) to physical states. Accordingly

− − c Φu ⋆ The first relation implies c0 Φt = Φu and this together with (2.48) implies that Φu c0 Φt = ( ) . − − | ci | i c h | Φu| i − The second relation implies Φt = c0 Φu , which, using (2.58), gives us Φt = ( ) Φu c0 . This c | i − | i h | − h | result, together with Φ Φt = 1, agrees with the result derived from the first relation. h t | i

107 we take all of the B states to be physical and use the above to obtain

Φu 0= ( ) i Q Φ Φ , B , , B − | ui i· ti 1 ··· n g Xi 

Φp + σ(il, jk) ( ) Bi1 Bil , Φp Φp, Bj1 , , Bjk  − ··· g1 · ··· g2 gX1,g2 Xp    {il,jk}l,k e

Φui + ( ) Bi1 Bil , Φti Φui , Bj1 , , Bjk − ··· g1 · ··· g2 Xi   

Φui + ( ) Bi1 Bil , Φui Φti , Bj1 , , Bjk (9.10) − ··· g1 · ··· g2  Xi    1 ′ + ( )Φs ( )Φp Φ Φ , Φ , Φ , B , , B 2 −  − | pi· p s s 1 ··· n g−1 Xs Xp  e e Φu + ( ) i Φ Φ , Φ , Φ , B , , B − | ti i· ui s s 1 ··· n g−1 Xi  e Φu + ( ) i Φ Φ , Φ , Φ , B , , B . − | ui i· ti s s 1 ··· n g−1 Xi  e This equation is really three equations, that is, the coefficients of the physical , unphysical and trivial states should vanish. Let us pay attention now only to physical states. From the above equation we then find that for each physical state labeled by p′ we must have

Φp ′ 0= σ(il, jk) ( ) Φp , Bi1 Bil , Φp Φp, Bj1 , , Bjk  − ··· g1 · ··· g2 gX1,g2 Xp   {il,jk}l,k e Φu i ′ + ( ) Φp , Bi Bil , Φti Φui , Bj , , Bjk − 1 ··· g1 · 1 ··· g2 Xi   (9.11)

Φu i ′ + ( ) Φp , Bi Bil , Φui Φti , Bj , , Bjk − 1 ··· g1 · 1 ··· g2  Xi   ′ 1 Φs + ( ) Φ ′ , Φ , Φ , B , , B . 2 − p s s 1 ··· n g−1 Xs  e Let us now consider the case when the multilinear functions use the full moduli spaces , Mg,n and become the string amplitudes. BRST decoupling of trivial states in Eqn.(9.4) implies that the second and third term in the above equation vanish. Therefore, if we work at genus zero, when the last term is not present, we have

Φp 0= σ(i , j )( ) Φ ′ , B B , Φ Φ , B , , B , (9.12) l k − p i1 ··· il p 0 · p j1 ··· jk 0 p,{il,jXk}l≥1,k≥2   l+k=n e which is the form of the Ward identities derived in Ref. [17] in the context of 2D string theory. They involve only physical states. For the case of higher genus we must still do some work since

108 the sum over states s in the quantum term of (9.11) involves physical, as well as unphysical and trivial states. Consider this term and expand the sum over states as usual

′ 1 Φs 1 Φp ( ) Φ ′ , Φ , Φ , B , , B = ( ) Φ ′ , Φ , Φ , B , , B 2 − p s s 1 ··· n g−1 2 − p p p 1 ··· n g−1 Xs  Xp  e e 1 Φt + ( ) i Φ ′ , Φ , Φ , B , , B 2 − p ti − ui 1 ··· n g−1 Xi 

1 Φu + ( ) i Φ ′ , Φ , Φ , B , , B . 2 − p ui ti 1 ··· n g−1 Xi  (9.13)

Since the statistics of Φui and Φti are opposite to each other the last two terms add (rather than cancel out!)

′ 1 Φs 1 Φp ( ) Φ ′ , Φ , Φ , B , , B = ( ) Φ ′ , Φ , Φ , B , , B 2 − p s s 1 ··· n g−1 2 − p p p 1 ··· n g−1 Xs  Xp  e e (9.14) Φu + ( ) i Φ ′ , Φ , Φ , B , , B . − p ui ui 1 ··· n g−1 Xi  e This is not surprising since we have not yet used the BRST properties of trivial and unphysical states. We will now show that indeed

Φu ( ) i Φ ′ , Φ , Φ , B , , B =0, (9.15) − p ui ui 1 ··· n g−1 Xi  e by use of the quartet mechanism of Kugo and Ojima [28]. The basic idea in this mechanism is that an unphysical state determines three other states, and the four together form a quartet. The unphysical state, acted by the BRST operator gives us a trivial state. These two states form a BRST doublet. Now, the unphysical state must have a tilde conjugate state, this conjugate state must be trivial, and therefore together with the state it arises from by action of Q forms another doublet. These two doublets form a quartet. This may be sketched in the following diagram Φ Qα u −→ u e Q Q (9.16)  x Φu  ( ) yQΦu αu − ←− e As we see, the unphysical field, via tilde conjugation gives us the trivial state Qαu, then clearly αu is unphysical. The nice property is that the tilde-conjugate of the unphysical state αu is the trivial state QΦu, with the sign given correctly in the diagram. As equations, the above diagram says that Φ = Q α α =( )Φu Q Φ . (9.17) | ui | ui → | ui − | ui e e 109 ⋆ Note that the states Φu and αu have the same statistics. Now, going back to Eqn. (9.15), we have to sum over unphysical states. We do this sum by pairing the unphysical states, and we will sum over the possible pairs (Φu, αu). We therefore get

Φu αu ( ) Φ ′ , Φ , Qα , B , , B + Φ ′ , α , ( ) QΦ , B , , B . (9.18) −  p u u 1 ··· n g−1 p u − u 1 ··· n g−1 pairsX  

It now follows from (9.5) and the graded-commutativity of the multilinear functions that

αu Φ ′ , α , ( ) QΦ , B , , B = Φ ′ , Qα , Φ , B , , B p u − u 1 ··· n g−1 − p u u 1 ··· n g−1   (9.19) = Φ ′ , Φ , Qα , B , , B , − p u u 1 ··· n g−1  and back in (9.18) we get the desired cancellation. It should be noted that for unphysical states whose unphysical partner is the same state (α Φ ), as may happen for states of u ∼ u ghost number two, the single contribution to the sum is vanishes by the same argument. Therefore we have verified that the unphysical and trivial states decouple from the sum and we finally have

Φp ′ 0= σ(il, jk) ( ) Φp , Bi1 Bil , Φp Φp, Bj1 , , Bjk − ··· g1 · ··· g2 gX1,g2 Xp i ,j l,k   { l k} e (9.20) 1 Φp + ( ) Φ ′ , Φ , Φ , B , , B . 2 − p p p 1 ··· n g−1 Xp  e This equation is the reduction of the main identity to physical states, and indeed all sums over states run only over the physical states of the theory (the semi-relative BRST cohomology classes). If we now define the new physical string product

B , B ( )Φp Φ Φ , B , , B , (9.21) 1 ··· n g ≡ − | pi· p 1 ··· n g   Xp  e where the sum only involves physical states, the previous identity reads

1 Φp ′ ′ 0= σ(il, jk) Φp , Bi1 Bil , Bj1 , , Bjk + ( ) Φp , Φp, Φp, B1, , Bn . ··· ··· g2 g1 2 − ··· g−1 gX1,g2    Xp  {il,jk}l,k e (9.22) As in Eqn.(4.11) which applies to the starting point for our present derivation of (9.22), we have that the sum over splittings runs over all l,k 0 with l + k = n 0, but with l 1, ≥ ≥ ≥ ⋆ A derivation of the signs is sketched for the convenience of the interested reader. The first relation in − − c c Φu Φu (9.17) gives Φu = c0 Q αu and this together with Φu Φu = ( ) gives us Φu c0 Q αu = ( ) . − | i | i h c | i Φu− h | | i − On the other hand the second relation in (9.17) gives αu = ( ) c0 Q Φu , and using (2.52) and (2.58), − − c Φu+1 | i c − Φ| u+1i one finds that αu = ( ) Φu Qc0 , which implies αu αu = ( ) Φu Qc0 αu = 1. This last h | − h | h | i − h | | i − − inner product agrees with the one following from the first relation in (9.17) after using Φu = Φu c0 b0 − h | h | to interchange the relative positions of Q and c0 .

110 when g = 0 and k 2 when g = 0. These latter restrictions are not necessary, however, 1 ≥ 2 ′ ′ since the l = 0 terms vanish due to Φp , Bj Bjk 0 = Φp , Q Bj Bjk = 0, and { 1 ··· g2 } h 1 ··· g2 i the k = 0 and k = 1 terms vanish because 0, and B QB= 0, respectively. · 0 ≡ 0 ≡ Let us now recover the master equation for physical states. We define now the on-shell string field Ψ which is simply the string field with the sum over states in the Hilbert space | oi of the conformal field theory restricted to the states in the semi-relative cohomology. Thus in defining fields and antifields we simply have

Ψ = Ψ + Ψ , Ψ = Φ ψp, Ψ = Φ ψ∗, (9.23) | oi | o−i | o+i | o−i | pi | o+i | pi p G(ΦXp)≤2 G(ΦXp)≤2 e We take for the master action, the same master action given in Eqn.(4.67), but where the string field is now the on-shell field defined above and the multilinear functions use the full moduli spaces. This will be called the on-shell master action. Since we only have physical fields, there is no genus zero quadratic term in the on-shell master action. Every term in the on-shell action is the string amplitude for the corresponding states. We will now argue that this on-shell master action satisfies the master equation.

Our first step is to get Eqn. (9.22) in a more convenient form. We take Bi =Ψo for all i, and moreover Φ ′ =Ψ . Then with l = n 1, k = n , and l + k = n 1, the above equation p o 1 − 2 − gives us

(n 1)! n1 n2 1 Φp n 0= − Ψo , Ψo + ( ) Φp, Φp, Ψo . (9.24) (n 1)!n ! g2 g1 2 − g−1 g1,g2 1 2 p n X,n −    X  1≥1 2≥0 e n1+n2=n≥1

Since this equation is entirely analogous to equation (4.63) we have that the above implies

n! n1 n2 1 Φp n 0= Ψo , Ψo g g + ( ) Φp, Φp, Ψo g−1. (9.25) n1!n2! 2 1 2 − gX1,g2    Xp  n1≥1,n2≥0 e n1+n2=n≥1

which is the analog of Eqn.(4.66). Since this equation was the basic ingredient in the verifi- cation of the master equation in the general case, we will now just go over the steps involved and see that everything goes through in the present conditions. The derivation begins in ′ Eqn.(4.72), and wherever we see s we must remember the sum is only over physical states. The derivation goes through untilP Eqn.(4.76), where it is clear that we can indeed replace the ′ restricted sum over physical states by the complete sum s. Then we can simply use the derivation up to Eqn.(4.80). The analysis of the second termP of the BV equation simply goes through, using the restricted sums. This shows that indeed the BV master equation is satisfied by the on-shell master action and concludes our derivation. In Ref.[17] the action reproducing the three point couplings of the physical states represents (the restriction to ghost number two states of) the cubic term in the on-shell master action

111 for two-dimensional closed string field theory. The physical states in two-dimensional closed string theory can be described in terms of differential forms on a three dimensional space and their couplings in terms of in that space. The significance of the on-shell master action suggests that it should be of interest to derive its complete form for two-dimensional closed string theory.

Acknowledgements: I am indebted to E. Witten for discussions that contributed much to the understanding of the algebraic structure of the theory. I am grateful to A. Giveon and J. Stasheff for their useful comments on the paper. I wish to thank P. Nelson for discussions on the operator formalism, to K. Ranganathan for discussions on Schiffer variations, and to M. Wolf for discussions on the minimal area problem. Useful conversations with M. Kontsevich, A. S. Schwarz and W. Thurston are gratefully acknowledged.

REFERENCES

1. M. Saadi and B. Zwiebach, ‘Closed string field theory from polyhedra’, Ann. Phys. 192 (1989) 213. 2. T. Kugo, H. Kunitomo and K. Suehiro, ‘Non-polynomial closed string field theory’, Phys. Lett. 226B (1989) 48. 3. T. Kugo and K. Suehiro, ‘Nonpolynomial closed string field theory: action and gauge invariance’, Nucl. Phys. B337 (1990) 434. 4. M. Kaku, ‘Geometrical derivarion of string field theory from first principles: closed strings and modular invariance. Phys. Rev. D38 (1988) 3052; M. Kaku and J. Lykken, ‘Modular Invariant closed string field theory’, Phys. Rev. D38 (1988) 3067. 5. B. Zwiebach, ‘Quantum closed strings from minimal area’, Mod. Phys. Lett. A5 (1990) 2753. 6. B. Zwiebach, ‘Recursion Relations in Closed String Field Theory’, Proceedings of the “Strings 90” Superstring Workshop. Eds. R. Arnowitt, et.al. (World Scientific, 1991) pp 266-275. 7. H. Sonoda and B. Zwiebach, ‘Closed string field theory loops with symmetric factorizable quadratic differentials’, Nucl. Phys. B331 (1990) 592. 8. B. Zwiebach, ‘How covariant closed string theory solves a minimal area problem’, Comm. Math. Phys. 136 (1991) 83 ; ‘Consistency of closed string polyhedra from minimal area’, Phys. Lett. B241 (1990) 343. 9. W. Siegel, ‘Introduction to String Field Theory’ (World Scientific, Singapore, 1988). 10. E. Witten, ‘Noncommutative geometry and string field theory’, Nucl. Phys. B268 (1986) 253. 11. I. A. Batalin and G. A. Vilkovisky, ‘Quantization of gauge theories with linearly depen- dent generators’, Phys. Rev. D28 (1983) 2567.

112 12. M. Henneaux, ‘Lectures on the antifield-BRST formalism for gauge theories’, Proc. of the XXI GIFT Meeting. 13. M. Henneaux and C. Teitelboim, ‘Quantization of gauge systems’, Princeton University Press, Princeton, New Jersey, 1992, to be published. 14. A. Sen, Nucl. Phys. B345 (1990) 551; B347 (1990) 270. 15. K. Ranganathan, ‘A Criterion for flatness in minimal area metrics’, MIT-preprint MIT- CTP-1945, to appear in Comm. Math. Phys. 16. M. Wolf and B. Zwiebach, ‘The plumbing of minimal area surfaces’, IASSNS-92/11, submitted to Comm. Math. Phys. [email protected]. #9202062. 17. E. Witten and B. Zwiebach, ‘Algebraic Structures and Differential Geometry in 2D String Theory’, IASSNS-HEP-92/4, to appear in Nucl. Phys. B. [email protected]. #9201056 18. E. Verlinde, ‘The Master Equation of 2D String Theory’ IASSNS-HEP-92/5, to appear in Nucl. Phys. B. [email protected]. #9202021. 19. A. Sen, ‘Some applications of string field theory’, TIFR/TH/91-39, Talk at the ‘Strings and Symmetries, 1991’, conference held at Stony Brook, May 25-30, 1991. [email protected]. #9109022. 20. J. Stasheff, ‘Towards a closed string field theory: topology and convolution algebra’, University of North Carolina preprint, UNC-MATH-90/1, to appear. 21. J. Stasheff, ‘Homotopy associativity of H-spaces, II.’, Trans. Amer. Math. Soc., 108 (1963) 293; ‘H-Spaces from a homotopy point of view’, Lecture Notes in Mathematics 161 Springer Verlag, 1970. 22. E. Witten, ‘Chern-Simons gauge theory as a string theory’, IASSNS-92/38, June 1992. 23. M. Kontsevich, ‘The Cohomology of Graphs’, to appear. 24. C. B. Thorn, ‘Perturbation Theory for Quantized String Fields’, Nucl. Phys. B287 (1987) 61. 25. C. B. Thorn, ‘String field theory’, Phys. Rep. 174 (1989) 1. 26. M. Bochicchio, ‘Gauge fixing for the field theory of the bosonic string’, Phys. Lett. B193 (1987) 31. 27. A. Sen, ‘Equations of motion in non-polynomial closed string field theory and conformal invariance of two dimensional field theories’, Phys. Lett. B241 (1990) 350. 28. T. Kugo and I. Ojima, ‘Local Covariant Operator Formalism of Non-Abelian Gauge Theories and Quark Confinement Problem’, Suppl. Prog. Theor. Phys. 66 (1979) 1. 29. K. Strebel, “Quadratic Differentials”, Springer-Verlag (1984) 30. K. Ranganathan, ‘On the background independence of overlap string interactions’, MIT preprint, to appear.

113 31. T. Kugo and B. Zwiebach, ‘Target space duality as a symmetry of string field the- ory’, YITP/K-961, IASSNS-HEP-92/3, January 1992, to appear in Prog. Theor. Phys. [email protected]. #9201040. 32. R. Saroja and A. Sen, ‘Picture changing operators in closed fermionic string field theory’, TIFR-TH-92/14, February 1992. [email protected]. #9202087. 33. S. Shenker, unpublished. See also S. Shenker, ‘The strength of nonperturbative effects in string theory’. Proc. Cargese Workshop on Random Surfaces, Quantum Gravity and Strings, 1990. 34. A. A. Belavin, A. M. Polyakov, and A. B. Zamolodchikov, ‘Infinite conformal symmetry in two-dimensional quantum field theory’, Nucl. Phys. B241 (1984) 333. 35. C. Schubert, ‘The finite gauge transformations in closed string field theory’, MIT preprint, MIT-CTP-1977, May 1991, to appear in Mod. Phys. Lett. A. 36. L. Alvarez-Gaume, C. Gomez, G. Moore and C. Vafa, Nucl. Phys. B303 (1988) 455; C. Vafa, Phys. Lett. B190 (1987) 47. 37. P. Nelson, Phys. Rev. Lett. 62 (1989) 993; H. S. La and P. Nelson, Nucl. Phys. B332 (1990) 83; J. Distler and P. Nelson, Comm. Math. Phys. 138 (1991) 273. 38. A. LeClair, M. E. Peskin, and C. R. Preitschopf, Nucl. Phys. B317 (1989) 411. 39. H. Hata, ‘BRS invariance and unitarity in closed string field theory’, Nucl. Phys. B329 (1990) 698; ‘Construction of the quantum action for path-integral quantization of string field theory’, Nucl. Phys. B339 (1990) 663. 40. B. Zwiebach, ‘Quantum Open string theory with manifest closed string factorization’, Phys. Lett. 256B (1991) 22; ‘Interpolating string field theories’, Mod. Phys. Lett. A7 (1992) 1079.

41. E. Getzler and J. D. S. Jones, ‘A∞ algebras and the cyclic bar complex’, Ill. Jour. Math. 34 (1990) 256. 42. W. Thurston, private communication (1991). 43. R. Brustein and S. P. De Alwis, Nucl. Phys. B352 (1991) 451. 44. R. Brustein and K. Roland, ‘Space-Time versus World-Sheet Renormalization Group Equation in String Theory’, Nucl. Phys. B372 (1992) 201. 45. S. Giddings and E. Martinec, ‘Conformal geometry and string field theory’, Nucl. Phys. B278 (1986) 91. 46. H. Sonoda and B. Zwiebach, ‘Covariant closed string theory cannot be cubic’, Nucl. Phys. B336 (1990) 185. 47. G. Zemba and B. Zwiebach, ‘Tadpole graph in covariant closed string field theory’, J. Math. Phys. 30 (1989) 2388. 48. M. Schiffer and D. C. Spencer, “Functionals of Finite Riemann surfaces”, Princeton University Press, Princeton, New Jersey, 1954.

114 49. F. P. Gardiner, “Schiffer’s Interior Variation and quasiconformal mappings”. Duke. Math. Jour. 42 (1975) 371. 50. S. Nag, ‘Schiffer variation of complex structure and coordinates for Teichmuller spaces’, Proc. Indian Acad. Sci. (Math. Sci.) 94 (1985) 111. 51. H. Sonoda, ‘Sewing Conformal Field Theories: I,II’, Nucl. Phys. B311 (1988) 401,417. 52. M. Spivak, ‘A comprehensive introduction to differential geometry’, Publish or Perish Press, Berkeley, 1979. 53. I. Klebanov, ‘Ward Identities in Two-Dimensional String Theory’, PUPT-1302, Decem- ber 1991

FIGURE CAPTIONS

1) Consider the family of metrics ρ with t [0, 1]. The area function (ρ ) is a strictly t ∈ A t convex function of t. This can be used to show the uniqueness of the minimal area metric. 2) Two saturating geodesics and intersect at a point p where the metric is smooth. C1 C2 1 One can show their intersecting elsewhere (as in the figure) leads to a contradiction. 3) (a) A minimal area metric must be isometric to a semiinfinite cylinder of circumference 2π for some neighborhood of each puncture. This cylinder must end somewhere: the boundary of the maximal region with saturating geodesics homotopic to the puncture is the curve . A curve is the saturating geodesic in the cylinder a distance l away C0 Cl from . The coordinate curves will be taken to be the curves. (b) The property of C0 Cπ amputation states that if we amputate the semiinfinite cylinder along for l > 0, the Cl metric of minimal area on the cut surface is simply the restriction of the original metric on the complete surface. 4) This figure is used to prove the cutting theorem. Shown in (a) is a surface which has R a foliation F of height 2π +2δ. This surface is cut along the middle curve to obtain C the cut surface shown in (b). In (c) we show a typical curve crossing . Rc C 5) This figure is used for the proof that a Schiffer vector that extends holomorphically away from the puncture does not change any data. In (a) we show the surface broken up R into a disk D and the rest. In (b) we show the surface ′ obtained by variation. In (c) R we show once more the surface broken up in a different way. R

115